Holmes MH Introduction To Differential Equations
Holmes MH Introduction To Differential Equations
HOLMES
Introduction
to Differential
Equations 2e
Introduction to Differential
Equations
Mark H. Holmes
$PQZSJHIUªCZ.BSL))PMNFT
"MMSJHIUTSFTFSWFE
1SJOUFEJOUIF6OJUFE4UBUFTPG"NFSJDB
*4#/
/PQBSUPGUIJTCPPLNBZCFSFQSJOUFEJOBOZNBOOFSXJUIPVUXSJUUFOQFSNJTTJPO
GSPNUIFQVCMJTIFS
7FOUVSF%SJWF 4VJUF
"OO"SCPS .*
XXXYBOFEVDPN
Contents
Preface v
1 Introduction 1
1.1 Terminology for Differential Equations . . . . . . . . . 2
1.2 Solutions and Non-Solutions of Differential Equations . 3
2 First-Order Equations 7
2.1 Separable Equations . . . . . . . . . . . . . . . . . . . . 7
2.1.1 General Version . . . . . . . . . . . . . . . 8
2.2 Integrating Factor . . . . . . . . . . . . . . . . . . . . . 14
2.2.1 General and Particular Solutions . . . . . . 16
2.2.2 Interesting But Tangentially Useful Topics 17
2.3 Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.3.1 Mixing . . . . . . . . . . . . . . . . . . . . 20
2.3.2 Newton’s Second Law . . . . . . . . . . . . 23
2.3.3 Logistic Growth or Decay . . . . . . . . . . 24
2.3.4 Newton’s Law of Cooling . . . . . . . . . . 26
2.4 Steady States and Stability . . . . . . . . . . . . . . . . 33
2.4.1 General Version . . . . . . . . . . . . . . . 35
2.4.2 Sketching the Solution . . . . . . . . . . . . 36
2.4.3 Parting Comments . . . . . . . . . . . . . . 38
i
ii Contents
4 Linear Systems 79
4.1 Linear Systems . . . . . . . . . . . . . . . . . . . . . . 80
4.1.1 Example: Transforming to System Form . . 81
4.1.2 General Version . . . . . . . . . . . . . . . 82
4.2 General Solution of a Homogeneous Equation . . . . . 84
4.3 Review of Eigenvalue Problems . . . . . . . . . . . . . 85
4.4 Solving a Homogeneous Equation . . . . . . . . . . . . 91
4.4.1 Complex-Valued Eigenvalues . . . . . . . . 92
4.4.2 Defective Matrix . . . . . . . . . . . . . . . 93
4.5 Summary for Solving a Homogeneous Equation . . . . 94
4.6 Phase Plane . . . . . . . . . . . . . . . . . . . . . . . . 99
4.6.1 Examples . . . . . . . . . . . . . . . . . . . 99
4.6.2 Connection with an IVP . . . . . . . . . . . 104
4.7 Stability . . . . . . . . . . . . . . . . . . . . . . . . . . 107
B Answers 225
Bibliography 236
Index 237
Preface
v
vi Preface
Mark H. Holmes
Department of Mathematical Sciences
Rensselaer Polytechnic Institute
May, 2020
Chapter 1
Introduction
We begin with a question: why are most students who are majoring
in engineering or science required to take an entire course dedicated to
something called differential equations?
dN
= −kN, for 0 < t, (1.1)
dt
where k is a positive constant. This is a differential equation for N .
Usually one knows the amount N0 of the isotope at the beginning, which
gives us the requirement that
N (0) = N0 . (1.2)
Introduction to Differential Equations, M. H. Holmes, 2020
1
2 Chapter 1. Introduction
Example 2: Mechanics
One of the biggest generators of differential equations is Newton’s second
law, which states that F = ma. Any situation, electrical, mechanical or
otherwise, involving non-static forces will almost inevitability result in
having to solve a differential equation. To illustrate, consider the simple
case of dropping an object off a building. If x(t) is the distance of the
object from the ground, then its velocity is v = x′ (t), and its acceleration
is a = x′′ (t). If the forces on the object are gravity Fg = −mg, and air
resistance Fr = −cv, then F = Fg +Fr . Together, these expressions result
in the following differential equation for x(t):
d2 x dx
m 2
= −mg − c . (1.3)
dt dt
The differential equations in (1.1) and (1.3) have a few things in com-
mon, such as there is one independent variable, t, and one dependent vari-
able, N and x. There are also differences, and an example is that (1.3)
involves the second derivative and (1.1) only involves the first derivative.
It is important to be able to recognize these differences as they are often
used in this textbook to determine how to solve the problem.
d2 y dy ∂u ∂u
Example 1: 2
− 2 + 4ty = 0 Example 2: −2 = u3
dt dt ∂t ∂x
du
Example 3: =u+v+1
dt
dv
= −u + v
dt
Independent variable(s): These are usually time (t) and/or space (x).
Order: The order of the highest derivative in the equation (or equations).
Exercises
1. Show that the given function y(t) is a solution of the given differential
equation.
a) y = e2t − 1, y ′ = 2y + 2 e) y = et + 1, y ′′ + 2y ′ − 3y = −3
1
b) y = te−t , y ′ + y = e−t f) y = , y′ + y2 = 0
1+t
c) y = cos(3t), y ′′ = −9y g) y = tan 31 t + 1 , 3y ′ = 1 + y 2
a) y ′ = −2y f) y ′′ − 4y ′ + 4y = 0
b) 3y ′ = y g) y ′′ + y ′ + y = e−3t
c) y′ = y + 1 h) y ′′ − 3y ′ + y = 1
d) y ′′ + 4y ′ = 0 i) y ′ = −2y 3
e) 2y ′′ + 5y ′ − 3y = 0 j) y′ = y2
Exercises 5
a) y ′′ − 3y ′ + 2y = 0, c) y ′′ + y ′ = 0,
y1 = e2t , y2 = et y1 = e−t , y2 = 1
b) y ′′ − y ′ − 2y = 0, d) y ′′ + 2y ′ + 5y = 0,
y1 = e2t , y2 = e−t y1 = e−t cos(2t)
y2 = e−t sin(2t)
6. Both y1 (t) and y2 (t) are solutions of the given nonlinear differential
equation. Show that (i) y = c1 y1 (t) is not a solution unless c1 = 1,
and (ii) y = c1 y1 + c2 y2 is not a solution if c1 and c2 are both nonzero.
√
a) y ′ = t/(1 + y), b) y ′ = 1 + y,
y1 = −1 + t, y2 = −1 − t y1 = 14 t2 + t, y2 = 41 t2 + 2t + 3
7. Fill out the table on the next page. Assume that any constants in the
equation(s) are nonzero. Also, in the last column, the answer Inap-
plicable (IA) is possible. Reference for Schrödinger’s equation image:
Eigler [2020].
6
Radioactive decay
dy
dt
= −ry
Mass-Spring-Dashpot system
d2 y dy
m 2 + c + ky = sin t
dt dt
Pendulum equation
d2 θ g
dt2 ℓ
= − sin θ
Schrödinger’s equation
∂u ~2 ∂ 2 u
i~ +Vu
∂t 2m ∂x2
=−
Beam equation
∂4w ∂2w
+ 2 =P
∂x4 ∂t
Michaelis-Menten equations
dS
dt
= −k1 ES + k−1 (E0 − E)
dE
Chapter 1. Introduction
dt
= −k1 ES + k3 (E0 − E)
Chapter 2
First-Order Equations
dy
= f (t, y).
dt
There are no known analytical methods that can solve the general ver-
sion of this problem. Consequently, assumptions have to be made on
the function f (t, y) to be able to derive a solution. The two more useful
assumptions are that the function is separable or it is linear, and both
are considered in this chapter. The fact is, however, that for many real
world problems it is not possible to solve the differential equation by
hand. Consequently, the ability to determine the properties of the solu-
tion, without actually solving the problem, becomes essential. What this
entails is introduced in Section 2.4.
dy
= 3y 2 . (2.1)
dt
We are going to treat the derivative as if it were a fraction, and rewrite
the above equation as
dy
= 3dt. (2.2)
y2
So, the variables have been separated in the sense that all of the y terms
are on the left hand side, and the t terms are on the right. We now
Introduction to Differential Equations, M. H. Holmes, 2020
7
8 Chapter 2. First-Order Equations
y = 0. (2.5)
The method used to solve (2.1) is rather simple, but it contains the
questionable step of splitting the derivative to obtain (2.2). To explain
why this is possible, note that (2.1) can be written as y −2 dy
dt = 3. Using
d
the chain rule, this can be written as − dt (y −1 ) = 3. Integrating this
equation yields (2.3). So, the splitting the derivative step is effectively a
compact version of using the chain rule.
In theory, you carry out the above integrations, and then solve for y.
How difficult this might be depends on how complicated the y integral
is, and the examples that follow illustrate some of the complications that
can arise. It is also important to note that the above method requires
that G(y) 6= 0. Consequently, in addition to the solutions that come from
(2.8), you must include as solutions any constant that satisfies G(y) = 0.
dy 1
Z Z
− = dt.
y3 4
Integrating gives us
1 1
2
= t + c,
2y 4
which is rewritten as
2
y2 = .
t + 4c
From this we obtain the two solutions
r
2
y=± . (2.9)
t + 4c
dw x
Example 4: Solve = , where w(0) = −2.
dx 1+w
Answer: In this problem the independent variable is x and the
dependent variable is w. Separating variables, so (1 + w)dw = xdx,
and then integrating gives
Z Z
(1 + w)dw = xdx.
y
Example 5: Solve y ′ = − , where y(0) = 1.
1+y
Answer: Separating variables yields
1+y
dy = −dt.
y
Since (1 + y)/y = 1/y + 1, and y(0) > 0, then integrating we get
that
y + ln y = −t + c.
It is not possible to solve this for y as in the previous examples,
without resorting to more advanced mathematical methods. For
this reason, this is an example of what is called an implicit solu-
tion, and they are very common when solving nonlinear differential
equations. Even so, it is still possible to find c from the initial con-
dition. Substituting y = 1 and t = 0 into the above equation we get
2.1. Separable Equations 11
y + ln y = −t + 1. (2.10)
3t − 2c + 4
y= ,
t + 2c − 4
3t − c
y= .
t+c
Exercises
1. Find all of the solutions of the given differential equation.
3. Find the solution of the IVP. In these problems, the independent vari-
able is not t and the dependent variable is not y.
dq dz
a) = −7q 3 , q(0) = −1 e) (1+e−r ) +z 2 = 0, z(0) = 6
dr dr
dp dw
b) = −4p3 , p(0) = 0 f) 4 = τ 3 e−2w , w(0) = 0
dr dτ
dh dr
c) 3 = 2 + h, h(0) = 2 g) (θ + 1)3 = r2 , r(0) = 2
dτ dθ
dh dr 2θ
d) = h2 − 3h, h(0) = 2 h) = , r(0) = 0
dx dθ 1+r
Exercises 13
1 1+y
a) y ′ = 1 + , y(0) = 1 c) y ′ = , y(0) = 5
y 2+y
3 ey
b) y ′ = , y(0) = −1 d) y ′ = , y(0) = 2
1 + y4 1 + ey
y-axis
x-axis
d
(µy) = µ(t)y ′ (t) + µ′ (t)y(t). (2.13)
dt
The second is the Fundamental Theorem of Calculus, which states that if
d
(µy) = q(t),
dt
then Z t
µy = q(s)ds + c. (2.14)
0
The first step is the observation that the left hand side of (2.12) re-
sembles the right hand side of (2.13). To make it so they are exactly the
same we need to multiply the differential equation by µ(t), which gives us
µ′ = pµ. (2.16)
It will make the formula for the solution a bit simpler if we require
µ(0) = 1. (2.17)
The differential equation (2.16) is separable, and one finds that the solu-
tion that satisfies (2.17) is
Rt
p(r)dr
µ(t) = e 0 . (2.18)
2.2. Integrating Factor 15
With this choice for µ, the differential equation for y in (2.15) can be
written as
d
(µy) = µg. (2.19)
dt
From (2.14) we get that
Z t
µy = µ(s)g(s)ds + c,
0
The second special case arises for the homogeneous equation y ′ +p(t)y = 0.
Setting g = 0 in (2.20), gives us the solution
Rt
p(r)dr
y(t) = ce− 0 . (2.22)
From (2.18), the integrating factor is µ = e3t . So, since g(t) = e2t ,
then from (2.20),
Z t Z t
−3t 3s 2s −3t 5s
y(t) = e e e ds + c = e e ds + c .
0 0
16 Chapter 2. First-Order Equations
dh
Example 3: Solve − 4h = 2z, where h(0) = −1.
dz
Answer: In this problem the independent variable is z and the
dependent variable is h. The formula for the solution can still be
used, we just need to make the appropriate substitutions. Since
p = −4, then Z z Z z
p(r)dr = −4dr = −4z.
0 0
From (2.18), the integrating factor is µ = e−4z . So, since g(z) = 2z,
then from (2.21),
Z z
4z −4s
h(z) = e 2se ds − 1
0
1 7
= − (4z + 1) − e4z .
8 8
is Z t
1
y(t) = µ(s)g(s)ds + c . (2.25)
µ(t) 0
2.2. Integrating Factor 17
Any, and all, solutions of (2.24) are included in this formula, and for this
reason (2.25) is said to be the general solution.
A useful observation about (2.25) is that it can be written as
where
t
1
Z
yp (t) = µ(s)g(s)ds, (2.27)
µ(t) 0
and Rt
c
yh (t) = = ce− 0 p(r)dr . (2.28)
µ(t)
The formulas for yp and yh are not important here. What is important is
that yp is a solution of the differential equation (2.24). It does not contain
the arbitrary constant, and for this reason it is said to be a particular
solution. In contrast, the function yh (t), which contains an arbitrary
constant, is a solution of the differential equation
y ′ + p(t)y = 0. (2.29)
y ′ + ay = g(t), (2.30)
For those who have taken a course in linear algebra, there is a connec-
tion between that subject and linear differential equations that is worth
knowing about. To explain, a central problem in linear algebra is to solve
Ax = b, where A is a m × n matrix. It’s possible to prove that if there
is a solution of this equation, then it has the form x = xp + xh , where
xp is a particular solution and xh is the general solution of the associated
homogeneous equation Ax = 0. This is basically the same statement we
made for the solution of the linear differential equation (2.24). The key
property these equations have in common is that they are both linear. A
consequence of this is that the principle of superposition can be used (see
page 5) when solving the associated homogeneous equation. We will make
use of this fact in every chapter of this textbook, except for Chapter 5.
This illustrates the beauty, and profundity, of mathematical abstraction.
Namely, it is possible to make rather significant conclusions about the so-
lution of an equation, irrespective of whether it is algebraic or differential,
simply from the basic properties these equations have in common.
Exercises
1. Find the general solution of the given differential equation.
Exercises 19
3. Find the solution of the IVP. In these problems, the independent vari-
able is not t and the dependent variable is not y.
dq dz
a) + 2q = 4 , q(0) = −1 d) = 4z + 1 + τ , z(0) = 0
dz dτ
dp dh
b) + 4p = −x, p(0) = 0 e) (x + 7) + h = −1, h(0) = 2
dx dx
dw dh
c) 2 − w = e2τ , w(0) = 0 f) (5z+1) +5h = 3, h(0) = −1
dτ dz
a) y ′ − 2y = 6 c) 7y ′ − y = e2t + 3
b) y ′ + y = 3e−t d) y ′ + 2ty = 1
5. A Maxwell viscoelastic material is one for which the stress T (t) and
the strain-rate r(t) satisfy
dT
T +τ = κr,
dt
where τ and κ are positive constants. By solving this equation for T ,
and assuming T0 = T (0), show that
t
κ
Z
T = T0 e −t/τ
+ e(s−t)/τ r(s)ds.
τ 0
2.3 Modeling
The principal objective of the examples to follow is to show how a dif-
ferential equation is the mathematical consequence of the assumptions
about a physical system.
2.3.1 Mixing
Typical mixing problems involve a continuously stirred tank, as illustrated
in Figure 2.2. As an example, suppose that water, containing salt, is
flowing into a well-stirred tank. At the same time, the mixture in the
tank is flowing out. The goal is to determine how much salt is in the tank
as a function of t.
Rin : This is the rate that salt is flowing into the tank. If the incoming
volumetric flow rate is Fin , and cin is the concentration of salt in the
incoming water, then Rin = cin Fin .
Rout : This is the rate that salt is flowing out of the tank. If the outgoing
volumetric flow rate is Fout , then Rout = cFout .
2.3. Modeling 21
If the initial amount of salt in the tank is Q0 , then the resulting IVP for
Q is:
dQ
= Rin − Rout ,
dt
Q(0) = Q0 .
Example 1
Suppose that salt water, containing 1/2 lbs of salt per gal, is poured into
a tank at 2 gal/min. Also, the water flows out of the tank at the same
rate. If the tank starts out with 100 gal of water, with 10 lbs of salt per
gal, find a formula for the total amount of salt in the tank.
Setup
Note that because of the way the variables have been defined, Q is mea-
sured in pounds and t is measured in minutes.
Solution
Example 2
Salt water, containing 3 lbs of salt per gal, flows into a 50 gal drum at 2
gal/sec. If the drum initially contains 10 gal of pure water, find a formula
for Q as a function of t.
Comments about this problem: There is no outflow, so the volume of
water will increase. However, it’s a 50 gal drum, so eventually it will fill
and start running over. When this occurs there is outflow, at a rate equal
to the incoming rate. To account for this, the problem needs to be split
into two phases, one where the volume is increasing, and the second when
it is a constant.
Solution
The solution is Q(t) = 6t. Also, the volume of water in the tank is
V = 10 + 2t. So, this solution for Q holds for V ≤ 50, which means that
t ≤ 20.
Phase 2: As before, Rin = 6. For the outflow, the rate is 2 gal/sec and
the concentration in the outflow is Q/50. This means that Rout = Q/25.
Now, this phase starts at t = 20, and the amount of salt in the tank at
the start is 120 (this comes from the solution for Phase 1). This means
that the problem to solve is
dQ 1
= 6 − Q, for 20 < t,
dt 25
Q(20) = 120.
What is different about this problem is the time interval, which is not the
usual 0 ≤ t. However, this does not interfere with our solution methods,
and the solution can be found using an integrating factor or separation of
variables. One finds that the general solution of the differential equation
is
Q(t) = 150 + Ae−t/25 .
From the requirement that Q(20) = 120 it follows that A = −30e4/5 .
dv
m = F. (2.33)
dt
What sort of differential equation this might be depends on how F de-
pends on v. Once (2.33) is solved for v, then the position is determined
by integrating the equation
dx
= v. (2.34)
dt
Typically, the initial velocity v(0) and initial position x(0) are given,
and these are used to determine the integration constants obtained when
solving the problem.
Vertical Motion
The object is assumed to be moving vertically, either up or down. In this
case, x(t) is the distance of the object from the ground. It is also assumed
that it is acted on by gravity, Fg , and a drag force, Fr . Consequently, the
total force is F = Fg + Fr . As for what these forces are:
Drag force: As long as the object is not moving very fast, the drag is
proportional to the velocity (see Exercise 10). In this case, Fr = −cv,
where c is a positive constant (the minus sign is because the force is in
the opposite direction to the direction of motion).
Units and Values: In the exercises, the value to use for g is usually stated.
If it is not given, then you should leave g unevaluated. Whatever value is
used, it is only approximate. If a more physically realistic value is needed,
then you should probably use the Somigliana equation. Finally, weight
is a force, so for an object that weighs w lbs, its mass can be determined
from the equation w = mg.
Question 1: What is the resulting IVP for v, and what problem must be
solved to find x?
Answer: Since F = Fg + Fr = −mg − cv, where m = 2 and c = 1/2,
then from (2.33) the differential equation is
dv 1
= −10 − v. (2.35)
dt 4
Since the object is dropped from rest, then the initial condition is
v(0) = 0. Once v is known, then x is found by integrating (2.34),
and using the fact that x(0) = 1000. Also, note that v is measured
in meters per second, t is measured in seconds, and x in meters.
Question 2: What is the solution of the IVP, and the resulting solution
for x?
Answer: Using the integrating factor solution (2.21), it is found
that the general solution is v = −40 + ce−t/4 . Applying the initial
condition we get that
vT = lim v(t).
t→∞
′ P
P =r 1− P, (2.37)
N
which is known as the logistic equation. This nonlinear equation can be
solved using separation of variables, and partial fractions. Doing this, in
the case of when 0 < P < N ,
N dP
= rdt (2.38)
(N − P )P
⇒
1 1
Z
+ dP = rt + c
P N −P
⇒
P
ln = rt + c
N −P
⇒
P
= ert+c .
N −P
From this, we get
P = (N − P )c̄ert , (2.39)
where c̄ = ec is a positive constant. Doing the same thing for the case of
when N < P , one again gets (2.39) except that c̄ is a negative constant.
Moreover, for the divide by zero case of when P = 0, you get (2.39) but
c̄ = 0. In other words, except for when P = N , (2.39) holds with the
understanding that c̄ is an arbitrary constant. Solving (2.39) for P yields
N c̄ert
P = , (2.40)
1 + c̄ert
where c̄ is an arbitrary constant. If P (0) = P0 , and if P0 6= N , then
one finds that c̄ = P0 /(N − P0 ). When P (0) = N , this is a divide by
zero situation in (2.38), and the resulting solution is just the constant
P (t) = N .
The solution we have derived in (2.40) is known as the logistic function
or the logistic curve. When plotted for −∞ < t < ∞ it has a S, or
26 Chapter 2. First-Order Equations
P(0)
P-axis
0
0
t-axis
Figure 2.3. The logistic function (2.40), for −∞ < t < ∞, in the case of
when P (0) < N .
If the initial temperature of the object is T0 , then the resulting IVP for
T is:
dT
= −k(T − Ta ), (2.41)
dt
T (0) = T0 . (2.42)
This problem can be solved using the integrating factor solution (2.21), or
by using separation of variables. It is found that T = Ta + (T0 − Ta )e−kt .
T = 70 + 130e−kt . (2.43)
Question 3: When should you start drinking the coffee (according to the
National Coffee Association)?
Answer: The time when T = 140 occurs when 140 = 70 + 130e−kt ,
from which one finds that
ln(13/7)
t=2 min. (2.44)
ln(13/11)
Question 4: What is the computed value for the answer for Question 3?
Answer: It is t ≈ 7.4 minutes.
Reality Check : The models that are considered here are used to illustrate
how, and where, differential equations arise. As with all models, simplify-
ing assumptions are made to obtain the resulting mathematical problem.
Many of these assumptions are not considered or accounted for in our
examples, and the same is true for the exercises. As a case in point, New-
ton’s Law of Cooling is usually limited to cases of when |T − Ta | is not
very large, and its applicability depends on whether the heat flow is due
to conduction, convection, or radiation. Said another way, if you want to
impress your family at Thanksgiving by using the solution of the cook-
ing a turkey exercise (see below), just make sure to check on the turkey
temperature regularly to make sure your predictions are correct.
Exercises
In answering
√ the following questions, do not numerically evaluate numbers
2
such as 2, π/3, e , ln(4/3), etc. The exception to this is when the
question explicitly asks you to compute the answer.
So, for radioactive dating you know N0 , as well as the current value
of N . Explain how knowing N0 , N , and k can be used to determine
t (which is the time that has passed since the organism died).
d) Measurements in 1991 determined that the amount of carbon-14
in the Temple Scroll, which is one of the Dead Sea scrolls found
at Qumran, to be 186.18. The amount in living organisms is 238.
Determine (i.e., compute) what two years the scroll could have been
written in. Note that in the BC/AD system there is no year zero,
so it goes from 1 BC to 1 AD.
Comments: In this problem, the amount of carbon-14 refers to the
amount relative to carbon-12. Also, the organism is the parchment
from the scroll, and the testing is described in Bonani et al. [1992].
Mixing
3. A tank contains 100 L of salt water with a concentration of 2 g/L. To
flush the salt out, pure water is poured in at 4 L/min, and the mixture
in the tank flows out at the same rate.
a) What is the resulting IVP for the total amount Q(t) of salt in the
tank?
b) Solve the IVP determined in part (a).
c) How long does it take until the amount of salt in the tank is 1% of
its original amount?
4. A tank contains 20 L of fresh water. Suppose water, containing 14 g/L
of salt, starts to flow into the tank at 2 L/min, and the well-stirred
mixture flows out at the same rate.
a) What is the resulting IVP for the amount Q(t) of salt in the tank?
b) Solve the IVP determined in part (a).
c) How much salt is in the tank after one hour?
5. Ten years ago, a factory started operation in a pristine valley. The
valley’s volume is 106 m3 . Each year the factory releases 105 m3 of
exhaust through its smoke stacks, and this exhaust contains 1000 kg of
pollutants. Assume that the well-mixed polluted air leaves the valley
at 105 m3 /yr.
a) What is the IVP for the amount of pollutant in the valley?
b) How much pollutant is in the valley now?
6. A small lake contains 60,000 gal of pure water. There is an inlet stream
of pure water into the lake, as well as an outlet stream, both flowing
at a rate of 100 gal/min. Suppose someone starts pouring water into
the lake at the rate of 10 gal/min that contains 5 lbs/gal of a chemical,
and they do this for 8 hours. While this happens the inlet stream of
pure water is unchanged, and the outflow rate from the lake remains
at 100 gal/min.
30 Chapter 2. First-Order Equations
a) What is the formula for the volume of the lake while the person is
pouring?
b) What IVP must be solved to determine the amount of the chemical
in the lake?
c) How much of the chemical is in the lake when the person stops
pouring?
d) Once the person stops pouring, what IVP must be solved to deter-
mine how much of the chemical is in the lake?
104
100
Drag Force
10-4
10-8 Experimental
F
r
10-12
10-6 10-4 10-2 100 102 104
Speed (m/sec)
Figure 2.4. Drag force on a smooth sphere as a function of the speed [Roos
and Willmarth, 1971, NASA, 2020]. The function Fr is used in Exercise 10.
11. It is often found that a population will grow exponentially if the popu-
lation is very small, and it will decrease exponentially if the population
is very large. A model for this is due to Beverton and Holt, and the
32 Chapter 2. First-Order Equations
equation to solve is
P
1− N
P′ =r P
P,
1+ N
where r and N are positive constants.
a) Assuming that P (0) = 12 N , solve the resulting IVP for P .
b) What is the limiting population P (∞) = limt→∞ P (t)?
12. The population of fish in a large lake can be modeled using the logistic
equation. If, in addition, the fish are caught at a constant rate h, the
equation for the population becomes
P
P′ =r 1− P − h,
N
where r and N are positive constants. In this problem take r = 4,
h = 750, and N = 1000. Also, P (0) = 1000.
a) Solve the IVP for P .
b) What is the limiting population? In other words, what is P (∞) =
limt→∞ P (t)?
Cooling or Heating
13. Suppose coffee has a temperature of 200◦ F when freshly poured, and
the room temperature is 72◦ F. In this exercise use Newton’s law of
cooling.
a) What IVP does the temperature of the coffee satisfy?
b) What is the solution of the IVP?
c) If the coffee cools to 136◦ F in five minutes, what is k?
d) When does the coffee reach a temperature of 150◦ F?
14. Redo the previous exercise but use the Dulong-Petit law of cooling.
15. To cook a turkey you are to put it into a 350◦ F oven, and cook it
until it reaches 165◦ F. In answering the following questions, assume
Newton’s law of cooling is used.
a) Suppose the turkey starts out at room temperature, which is 70◦ F.
What IVP does the temperature satisfy?
b) Suppose that after two hours in the oven, the temperature of the
turkey is 140◦ F. How much longer before it is done?
c) Suppose the turkey is taken from the refrigerator, which is set to
40◦ F, and put directly into the oven. How much longer does it take
to cook than when the turkey starts out at room temperature? The
value for k is the same as in part (b).
16. A homicide victim was discovered at 1 p.m. in a room that is kept at
70◦ F. When discovered, the temperature of the body was 90◦ F, and
one hour later it had dropped to 85◦ F.
2.4. Steady States and Stability 33
P ′ = f (P ), (2.45)
0
0 0.5 1 1.5 2
Figure 2.5. Solution of (2.45) and (2.46) in the case of when P (0) = 0.1,
and when P (0) = 4.5. The dashed red lines are the steady state values.
Unstable: Even though the initial value P (0) = 0.1 is close to the steady
state P = 0, the solution moves away from P = 0. This happens
because of f (P ). To explain, the function f (P ) is plotted in Figure
2.6. It shows that f (P ) > 0 for 0 < P < 3. So, in this interval
P ′ (t) > 0, and this means that P is increasing. Similarly, since
f (P ) < 0 for 3 < P , then P is decreasing in this interval. The
arrows in Figure 2.6 indicate the corresponding movement of P . The
conclusion we derive from the arrows is that if P (0) is anywhere in
0 < P < 3, then the solution will move away from the steady state
P = 0. Because of this, the steady state is said to be unstable.
Stable: The second conclusion we make from the arrows in Figure 2.6 is
that if P (0) is anywhere in 0 < P < 3, then the solution increases
-3
-6
0 1 2 3 4 5
Figure 2.6. The function f (P ) in (2.46). The two steady states are shown
by the reds dots. The arrows indicate the direction P moves in the respective interval.
2.4. Steady States and Stability 35
y ′ = f (y), (2.47)
The idea underlying asymptotic stability is that if y(0) is any point close
to Y , then
lim y(t) = Y. (2.48)
t→∞
Example 1: Find the steady states, and determine their stability, for
y ′ = y 2 − y − 6.
Steady States: The steady states are the points where f (y) = 0. From
Figure 2.7, this happens when y = −2, y = 1, and y = 3. These are
identified using red dots in the figure. From the graph it is evident
that f ′ (−2) < 0 and f ′ (3) < 0, and this means that y = −2 and
y = 3 are asymptotically stable. Similarly, since f ′ (1) > 0, then
y = 1 is unstable.
-4 -3 -2 -1 0 1 2 3 4
y(0) = 1.3: This point is located between two steady states, specifically,
1 < y(0) < 3. According to Figure 2.7, y(t) increases monotonically
in this interval, and asymptotically approaches y = 3. A curve with
these properties is shown in Figure 2.8.
y(0) = 0.8: In this case, the point is located between two steady states,
namely, −2 < y(0) < 1. From Figure 2.7, y(t) decreases monoton-
ically in this interval, and asymptotically approaches y = −2. A
curve with these properties is shown in Figure 2.8.
y(0) = −4: For this initial condition, according to the information in Fig-
ure 2.7, y(t) increases monotonically, and asymptotically approaches
y = −2. A curve with these properties is shown in Figure 2.8.
-1
-2
-3
-4
0
Figure 2.8. Solution curves obtained using the information in Figure 2.7.
The dashed red lines are the steady state values.
38 Chapter 2. First-Order Equations
mined. For example, nothing was said about how steep the curves are,
or whether they are concave up or down. It is possible to determine this
using Figure 2.7, but this level of analysis is not considered in this text.
Exercises
1. For each equation, verify that Y = 0 is a steady state. Determine if it
is unstable or asymptotically stable.
2. For each differential equation, find the steady states and determine if
they are asymptotically stable or unstable.
a) y ′ = y 2 + y − 2 c) y ′ = 4y − y 3 e) y ′ = y 4 − 3y 2 − 4
b) y ′ = y 3 − y d) y ′ = e−y − 2 f) y ′ = e2y − 4ey + 3
3. Sketch the solution curve for each of the given initial conditions.
a) y ′ = y 2 + y − 2 d) y ′ = e−y − 2
y(0) = −3; y(0) = 0 y(0) = 1; y(0) = 2
b) y ′ = y 3 − y e) y ′ = y 4 − 3y 2 − 4
y(0) = 3/4; y(0) = −1/4 y(0) = 1; y(0) = −3
c) y ′ = 4y − y 3 f) y ′ = e2y − 4ey + 3
y(0) = 1/2; y(0) = 3 y(0) = −1; y(0) = ln 2
Exercises 39
5. For the mixing problem given in (2.31), (2.32), sketch the solution
without using the formula for the solution. Make sure to explain how
you do this.
6. For the population problem in Exercise 11, on page 31, sketch the
solution without using the formula for the solution. Use this to answer
part (b) of that exercise.
7. For the drag on a sphere, as described in Exercise 10 on page 31, de-
termine the terminal velocity without solving the IVP. In other words,
answer part (c) using only part (a) of that exercise.
√
8. This problem concerns solving y ′ = − 1 + y, where y(0) = 0.
a) Using separation of variables, what is the solution? It helps to
note, from the differential equation, that y ′ (0) = −1.
b) Using the method outlined in Example 2, sketch the solution.
c) Sketch the solution you found in part (a). Assuming your sketch
in part (b) is correct, is there anything wrong with your solution
in part (a)? If so, how should it be modified? Does your sketch
from part (b) need to be modified as well?
9. The population of fish in a lake can be modeled using the logistic
equation. However, assuming that the fish are caught at a constant
rate h, the equation for the population becomes
P
P′ =r 1− P − h,
N
where r and N are positive constants.
a) Assuming that the loss due to fishing is small enough that 0 <
h < rN/4, find the two steady states for the equation. Label these
values as P1 and P2 , where P1 < P2 .
b) Determine whether P1 and P2 are unstable or asymptotically stable.
c) Letting f (P ) be the right hand side of the differential equation,
sketch f (P ) for 0 ≤ P < ∞. With this, answer the question in
Exercise 12(b) on page 32.
d) Assuming that P1 < P (0) < P2 , sketch the solution. Do the same
for the case of when P2 < P (0).
e) Sketch the solution if 0 < P (0) < P1 . In doing this remember that
P (t) can not be negative. Note that you will find that there is a
time te where extinction occurs, and the differential equation does
not apply to the fish population for te < t.
40 Chapter 2. First-Order Equations
a) y ′ = 1 + y 2 c) y ′ = (y − 4)(y − 3)(y − 1)
b) y ′ = y − 4 d) y ′ = (y − 2)(4 − y)
0
0 1 2 3 4 5 6 7 8
Figure 2.9. Plot used in Exercise 10. The starting point is y(0) = 2.
11. The following refer to the solution of (2.47), where f (y) is continuous.
Sketch a function f (y) so the stated conditions hold. Make sure to pro-
vide a short explanation of why your function satisfies the conditions
stated. If it is not possible to find such a function, explain why.
a) The solution is strictly monotone increasing for y < 0, is strictly
monotone decreasing for y > 0, and there are no steady states.
b) The only asymptotically stable steady state is Y = 0, and the only
unstable steady states are Y = −1 and Y = 1.
c) The only asymptotically stable steady state is Y = 0, and the only
unstable steady states are Y = 1 and Y = 2.
12. This problem concerns what is known as one-sided stability, or semi-
stability. The differential equation considered is
y ′ = 2(3 − y)2 .
Second-Order Linear
Equations
d2 y dy
2
+ p(t) + q(t)y = f (t), (3.1)
dt dt
where p(t), q(t), and f (t) are given. One of the reasons this equation gets
its own chapter is Newton’s second law, which, if you recall, is F = ma.
To explain, if y(t) is the displacement, then the acceleration is a = y ′′ ,
and this gives us the differential equation my ′′ = F . In this chapter we
are considering problems when F is a linear function of velocity y ′ and
displacement y. Later, in Chapter 5, we will consider equations where the
dependence is nonlinear. It is because of the connections with the second
law that f (t) in (3.1) is often referred to as the forcing function.
In the previous chapter, for first-order linear differential equations, we
very elegantly derived a formula for the general solution. This will not
happen for second-order equations. All of the methods derived in this
chapter are, in fact, just good, or educated, guesses on what the answer
is. There are non-guessing methods, and one example involves using a
Taylor series expansion of the solution. An illustration of how this is
done can be found in Exercise 8 on page 51.
To use a guessing approach, it becomes essential to know the math-
ematical requirements for what can be called a general solution. This is
where we begin.
41
42 Chapter 3. Second-Order Linear Equations
where y0 and y0′ are given numbers. Given that our solution methods
involve guessing, it is important that we know when to stop guessing and
conclude we have found the solution. This is why the next result is useful.
Existence and Uniqueness Theorem. If p(t), q(t), and f (t) are con-
tinuous for t ≥ 0, then there is exactly one smooth function y(t) that
satisfies (3.1) and (3.2).
d2 y dy
2
+ p(t) + q(t)y = 0. (3.3)
dt dt
We need to spend some time discussing what it means to be the general
solution of this equation. So, consider Exercise 5(a), in Section 1.2. As-
suming you did this exercise, you found that given solutions y1 = e2t and
y2 = et of y ′′ − 3y ′ + 2y = 0, then
is a solution for any value of c1 and c2 . What is important here is that this
is a general solution of the differential equation. As in the last chapter,
this means that any, and all, solutions of the differential equation are
included in this formula.
This gives rise to the question: what is required so a solution like the
one in (3.4) can be claimed to be a general solution? The key to answering
this is the uniqueness guaranteed by the above theorem. The specifics of
the analysis are not needed here. What is needed is the conclusion, which
is stated next.
3.2. General Solution of a Homogeneous Equation 43
To explain how the Wronskian comes into this problem, (3.5) must hold
on the interval 0 ≤ t < ∞. So, (3.5) can be differentiated, which gives
us the equation c1 y1′ + c2 y2′ = 0. This, along with (3.5), provides two
equations for c1 and c2 . It is not hard to show that if W (y1 , y2 ) 6= 0, then
the only solution to these two equations is c1 = c2 = 0. Consequently, y1
and y2 are independent. The rest of the proof, along with some additional
information, can be found in Exercises 5 and 6.
Example: Show that y = c1 e−3t + c2 et is a general solution of y ′′ + 2y ′ −
3y = 0.
Answer: In this case, y1 (t) = e−3t and y2 (t) = et . It is not
hard to show that they are solutions of the differential equation
(see Section 1.2). To check on independence, from (3.7), W =
e−3t et − (−3)e−3t et = 4e−2t . This is not zero, and so the functions
are independent. Therefore, y is a general solution.
44 Chapter 3. Second-Order Linear Equations
Exercises
1. Assuming b 6= 0, show that y1 = 1 and y2 = e−bt are independent
solutions of y ′′ + by ′ = 0.
2. Assuming ω 6= 0, show that y1 = cos(ωt) and y2 = sin(ωt) are inde-
pendent solutions of y ′′ + ω 2 y = 0.
3. Assuming ω 6= 0, show that y = c1 eωt + c2 e−ωt is a general solution of
y ′′ − ω 2 y = 0.
4. Show y = c1 e−αt + c2 te−αt is a general solution of y ′′ + 2αy ′ + α2 y = 0.
d
5. If y1 and y2 are solutions of (3.3), show that dt W + p(t)W = 0. Use
this to derive Abel’s formula, which is that
Rt
p(r)dr
W (y1 , y2 ) = ce− 0 ,
where c is a constant.
6. Let y1 = (t − 1)2 and y2 = −(t − 1)|t − 1|.
a) On the same axes, sketch y1 and y2 for 0 ≤ t < ∞.
b) Use (3.5) to show that y1 and y2 are linearly independent for 0 ≤
t < ∞. Hint: Consider t = 0 and t = 2.
c) Show that W (y1 , y2 ) = 0. Explain why this, together with the result
in part (b), does not contradict the Independence Test.
d2 y dy
+ b + cy = 0 (3.9)
dt2 dt
1 p
r= − b ± b2 − 4c . (3.11)
2
The case of when the roots are complex-valued requires a short introduc-
tion to complex variables, and so it is done last.
y = c1 ert + c2 tert .
46 Chapter 3. Second-Order Linear Equations
y = c 1 e r1 t + c 2 e r2 t
= c1 e(−2+3i)t + c2 e(−2−3i)t . (3.13)
ez = ex+iy = ex eiy
= ex cos y + i sin y .
(3.15)
y = c1 e(−2+3i)t + c2 e(−2−3i)t
= c1 e−2t cos 3t + i sin 3t + c2 e−2t cos 3t − i sin 3t
This last expression is the formula we are looking for. In this representa-
tion of the general solution, R and ϕ are arbitrary constants that satisfy
0 ≤ R, and 0 ≤ ϕ < 2π. The advantage of this form of the general
solution is that it is much easier to sketch the solution, and to determine
its basic properties. Its downside is that it can be a bit harder to find R
and ϕ from the initial conditions than the other two representations.
y = c 1 e r1 t + c 2 e r2 t (3.19)
y = d1 et cos(5t) + d2 et sin(5t).
As you might have noticed, in the above examples the formula for
the roots in (3.11) was not used. The reason is that it is much easier
to remember the way the characteristic equation is derived (by assuming
y = ert , etc) than by trying to remember the exact formula for the roots.
Exercises
π
1. Assuming that z1 = 1 + i, and z2 = e2+i 6 , find Re(z) and Im(z):
a) z = z1 − 8 c) z = z2 e) z = z1 z2
b) z = 2iz1 d) z = z1 + 4z2 f) z = (z2 )6
a) y ′′ + y ′ − 2y = 0 f) y ′′ − 6y ′ + 9y =0
b) 2y ′′ + 3y ′ − 2y = 0 g) 4y ′′ + 4y ′ + y =0
c) y ′′ + 3y ′ = 0 h) 4y ′′ + y = 0
d) 4y ′′ − y = 0 i) y ′′ − 2y ′ + 2y =0
e) y ′′ = 0 j) y ′′ + 2y ′ + 5y =0
Exercises 51
5. The roots of the characteristic equation are given. You are to find the
original differential equation (of the form given in (3.18)). If only one
value is given, that is the only root.
a) r = −1, 1 d) r = 0, 2 g) r = 2 ± 5i
b) r = 3, 5 e) r = 1 h) r = ±2i
c) r = ±2 f) r = 0
d2 y dy
+ b + cy = 0. (3.26)
dt2 dt
How to find the general solution of this has been discussed in some detail,
and formulas for the solution are given in Section 3.5.
So, what remains is to determine how to find a particular solution of
(3.24). As you should recall, a particular solution is any function that
satisfies the differential equation. Since any function will do, we are not
really picky on how this function is determined. In fact, our go-to method
is nothing more than guessing what a particular solution might be. For
those who prefer a more systematic approach, an alternative method is
derived in Section 3.9. The guessing method, what is called the method
of undetermined coefficients, is considered first.
y = 1 − t + c1 et + c2 e−2t ,
3.7. The Method of Undetermined Coefficients 53
y ′′ + y ′ + 2y = 5e3t . (3.28)
Since y ′ = −4A sin 4t + 4B cos 4t, and y ′′ = −16A cos 4t − 16B sin 4t,
then (3.29) requires that
The key observation coming from the last example is that if you be-
lieve a function needs to be included in the guess for yp , then all of its
derivatives must be included. So, looking at (3.29) you would expect that
cos(4t) needs to be part of the guess, which means you must also in-
clude sin(4t). You do not need to include 4 sin(4t), or −4 sin(4t), because
sin(4t) is multiplied by an arbitrary constant in the guess (3.30), and this
can account for any constant factors that might be generated by taking
a derivative.
There are two situations when this guessing approach runs into trou-
ble. One is easily fixable and this is demonstrated in the next example.
The other situation is not fixable, and the cause of the difficulty is illus-
trated in Example 7 below.
y ′′ + 4y = 3 cos 2t.
Answer: Given what happened in the last example, you would ex-
pect that to find a particular solution you would assume that
However, both cos 2t and sin 2t are solutions of the associated ho-
mogeneous equation. Because of this, the guess would give us that
y ′′ + 4y = 0, no matter what the values are for A and B. The fix is
to take the guess, and for the terms that are solutions of the associ-
ated homogeneous equation, multiply them by t. So, the modified
guess for this example would be
and
Equating the coefficients of the cos 2t and sin 2t terms we get that
−4A = 0 and 4B = 3. Therefore, A = 0, B = 43 , and a particular
solution is yp = 43 t sin 2t.
3.7. The Method of Undetermined Coefficients 55
Finally, since yh = c1 e3t + c2 e−2t , and the guess does not include e3t
or e−2t , then our guess is, indeed, complete.
eat eat
tn tn , tn−1 , · · · , 1
To find A and B we equated the coefficients of the cos 4t and sin 4t terms
in this equation. This can be done because these functions are linearly
independent, and this is explained below. This approach does not require
that you prove the functions are independent. Rather, if you think they
might be, and you then determine values for A and B so (3.33) is satisfied
based on this assumption, then you have found a particular solution.
The explanation of why linear independence can be used to determine
A and B starts with rewriting (3.33) as
c1 cos 4t + c2 sin 4t = 0, ∀t ≥ 0,
fact is that they are mostly unreadable. It is much easier to just remember
the rules used in formulate the guess, and the earlier examples should be
reviewed for the particulars.
However, some do find a table useful, and one is provided in Table
3.1. A few comments need to be made about what is listed. First, if f (t)
contains tn , as well as tn−1 , or tn−2 , or tn−2 , etc, then the guess for tn
is all that you need (see Example 4 above). Second, when solving (3.9),
if one the functions in the left column is a solution of the associated ho-
mogeneous differential equation the guess must be modified. The needed
modification was explained earlier (see Examples 2 and 6).
Exercises
1. Find the general solution of the given differential equation.
Exercises 59
a) y ′′ − y ′ − 6y = 6et i) y ′′ − 2y ′ + 5y = 5t2 + 4
b) y ′′ + 3y ′ + 2y = sin πt j) y ′′ + 2y ′ + 10y = 3et + 1
c) y ′′ + 4y ′ − 5y = 2t2 k) y ′′ − 3y ′ = t3 − 6
d) 5y ′′ − y ′ = e−t + 3 cos 2t l) 3y ′′ + y ′ − 2y = 3e−2t − et
e) 3y ′′ − 5y ′ − 2y = t3 − 2t m)y ′′ − 8y ′ + 17y = e4t sin t
f) 8y ′′ − 2y ′ − y = 4 + 5 sin 2t n) y ′′ − 5y ′ − 6y = −3 sin(t + 7)
g) y ′′ + 4y = tet o) y ′′ + 3y ′ + 2y = sin2 t
h) y ′′ − 5y ′ − 6y = 10t sin(3t) p) 4y ′′ + y ′ = sin(t) cos(t)
a) y ′′ + y ′ − 2y = t5 − t2 g) y ′′ − y ′ + 6y = e−t cos 3t
b) y ′′ + 4y = t cos t h) 4y ′′ − y = −2(t − 1)7
c) y ′′ + 4y = t + sin 2t i) y ′′ − 2y ′ + 2y = et−5 cos t
d) y ′′ − y ′ = 1 + sin t j) y ′′ + 4y = cos(2t + 3)
e) y ′′ + 3y ′ = 1 + e−3t k) y ′′ + 25y = −3 sin(5t + 7)
R1√
f) y ′′ + y ′ + 2y = t3 e−2t l) y ′′ + y ′ + y = 0 s cos(2t − s)ds
a) y ′ − 6y = 2et e) y ′ − 4y = t sin 2t
b) 3y ′ + 2y = sin πt f) y ′ − 6y = te−t + 2
c) y ′ + 3y = 2t g) 3y ′ + 2y = e−t cos(t)
d) 5y ′ − y = e−t + 3t h) y ′ − 2y = cos(2t + 5)
d2 y dy
2
+ b + cy = f (t), (3.36)
dt dt
it is assumed that
y = u1 (t)y1 (t) + u2 (t)y2 (t), (3.37)
where y1 and y2 are independent solutions of the associated homogeneous
equation. As you should notice, the guess (3.37) resembles the general
solution of the associated homogeneous equation. However, instead of
arbitrary constants, there are now unknown functions u1 and u2 . Our job
is to find these functions. Although it might not appear to be significant
right now, we are looking for a single function, yp , yet our guess contains
two unknown functions. This means that we have the option to pick
one of these two functions anyway we wish. We will use this option to
advantage to find yp .
Our task is simple in that (3.37) must be a solution of (3.36). So, in
preparation for substituting (3.37) into (3.36) note that
u′1 y1′ + u′2 y2′ + u1 (y1′′ + by1′ + cy1 ) + u2 (y2′′ + by2′ + cy2 ) = f. (3.39)
Using the fact that y1 and y2 are solutions of the associated homogeneous
equation, then the above equation reduces to
−W (y1 , y2 )u′1 = y2 f,
d2 y dy
2
+ b + cy = f (t),
dt dt
1 t√
Z
y = cos(2t) + s sin(2t) cos(2s) − cos(2t) sin(2s) ds
2 0
1 t√
Z
= cos(2t) + s sin 2(t − s) ds.
2 0
Exercises 63
Exercises
1. Using variation of parameters, find a particular solution of the given
differential equation.
a) 2y ′′ + 3y ′ − 2y = 25e−2t d) y ′′ + 3y ′ = t3/2 + 1
2
b) y ′′ − 2y ′ + 2y = 6 e) 5y ′′ − y ′ = 1+t
c) y ′′ + y ′ − 2y = 3 ln(1 + t) f) 4y ′′ − y = sin(1 + t2 )
2. Find the solution of the IVP where the differential equation comes
from the previous problem, and the initial conditions are y(0) = 1 and
y ′ (0) = 0.
3. The formula for a particular solution given in (3.43) applies to the more
general problem of solving y ′′ +p(t)y ′ +q(t)y = f (t). In this case, y1 and
y2 are independent solutions of the associated homogeneous equation
y ′′ + p(t)y ′ + q(t)y = 0. In the following, show that y1 and y2 satisfy
the associated homogeneous equation, and then determine a particular
solution of the inhomogeneous equation.
a) t2 y ′′ − t(t + 2)y ′ + (t + 2)y = 2t3 ; y1 (t) = t, y2 (t) = tet
b) ty ′′ − (t + 1)y ′ + y = t2 e2t ; y1 (t) = 1 + t, y2 (t) = et
c) t2 y ′′ − 3ty ′ + 4y = t5/2 ; y1 (t) = t2 , y2 (t) = t2 ln t
4. The Bessel equation of order p is t2 y ′′ + ty ′ + (t2 − p2 )y = 0. In this
problem assume that p = 21 .
√ √
a) Show that y1 = sin t/ t and y2 = cos t/ t are linearly independent
solutions for 0 < t < ∞.
b) Use the result from part (a), and the preamble in Exercise 3, to find
a particular solution of t2 y ′′ + ty ′ + (t2 − 1/4)y = t3/2 cos t.
64 Chapter 3. Second-Order Linear Equations
where f (t) is an external forcing function. In this equation, u(t) is the dis-
placement of the mass from its rest position, with positive in the upward
direction. The physical interpretation of the terms in this differential
equation, and the basic properties of the solution are described in the
following pages.
Figure 3.2. Left: The original spring. Middle: The situation after the mass
is attached, and at rest. Right: Displacement u(t) of the mass from its rest position.
equation
mu′′ + ku = 0. (3.49)
To find the general solution of (3.49), from the assumption that u = ert
the characteristic equation is found to be mr2 + k = 0. This produces the
roots r = ±ω0 i, where r
k
ω0 = . (3.50)
m
From (3.23), the general solution can be written as
where u0 and u′0 are given. To satisfy these, from (3.51), we get that
R cos(ϕ) = u0 , (3.53)
u′0
R sin(ϕ) = . (3.54)
ω0
Figure 3.3. The initial conditions as expressed in (3.53) and (3.54), located
using the black dot, and the value of R and ϕ.
66 Chapter 3. Second-Order Linear Equations
Finding R: Using the identity cos2 (ϕ) + sin2 (ϕ) = 1, it follows that
s
u ′ 2
0
R = u20 + . (3.55)
ω0
Finding ϕ: The value for ϕ depends on whether u0 and u′0 are positive or
negative, as illustrated in Figure 3.3. To compute ϕ, assuming that
u0 6= 0, the ratio of (3.54) with (3.53) yields
u′0
tan(ϕ) = .
ω0 u0
The principal value of the arctan function is denoted as Arctan, and it
satisfies − π2 < Arctan(z) < π2 . This is the value most calculators, or
programs like MATLAB, give when evaluating arctan(z). So, setting
z = u′0 /(ω0 u0 ),
Arctan(z)
if u0 > 0 and u′0 ≥ 0,
ϕ= Arctan(z) + π if u0 < 0, (3.56)
if u0 > 0 and u′0 < 0.
Arctan(z) + 2π
3.10.3 Damping
We will now include a damping mechanism. It is assumed that the damp-
ing force is proportional to the velocity. For the mass-spring system the
resistance is usually illustrated as a dashpot, as shown in Figure 3.1. Ir-
respective of exactly what mechanism is involved, the result is that the
damping force is Fd = −cv, where v = u′ is the velocity, and c is the
damping constant and it is non-negative. From the equation F = ma,
and the fact that F = Fs + Fg + Fd , the resulting differential equation is
lim u = 0. (3.60)
t→∞
68 Chapter 3. Second-Order Linear Equations
Critically damped: This means that the damping constant has just
that right value that c2 = 4mk. So, there is one root, and the resulting
general solution is
u = (c1 + c2 t)ert , (3.61)
where r = −c/(2m). So, as for the previous case, no matter what the
initial conditions, (3.60) holds.
0
3
0
3
-3
3
-3
0 2 4 6 8 10 12 14 16 18 20
3.10.4 Resonance
We now consider what happens when a simple harmonic oscillator is
forced periodically. The specific equation to solve is
50
u(t)
-50
0 50 100 150 200 250 300 350 400
10
u(t)
-10
0 50 100 150 200 250 300 350 400
t
Figure 3.7. Upper: Resonant solution given in (3.65) when ω√= ω0 = 1/3,
F = 1, m = 9 and k = 1. The red dashed lines correspond to ±F t/2 km. Lower:
Solution when a dashpot, with c = 1/4, is included.
3.10. Linear Oscillator 71
m(ω02 − ω 2 )A + cωB = F,
−cωA + m(ω02 − ω 2 )B = 0.
Figure 3.8. The amplitude (3.67) of the forced, but damped, oscillator, as a
function of the driving frequency. Note that ω0 is the natural frequency of the undamped
oscillator.
Now, suppose that for the flutter problem it is found experimentally that
the wings won’t break off if the amplitude of the oscillation satisfies R ≤
Rb . Based on our calculations, this means that the damping coefficient c
must be large enough that RM ≤ Rb .
Reality Check: The resonance phenomena considered here is not possi-
ble for the mass-spring system envisioned in Figure 3.2. As the oscillations
grow, as predicted by (3.65), they will eventually get to the point that the
mass will start banging up against the upper support. Presumably, as the
amplitude grows, our simple linear model is no longer valid, and a more
physically realistic, nonlinear, model is necessary. Even so, the simple
model is useful as it provides information about the onset of resonance.
Units and Values: In the exercises, the value to use for g is usually
stated. If it is not given, then you should leave g unevaluated. Whatever
value is used, it is only approximate. If a more physically realistic value is
needed, then you should probably use the Somigliana equation. Finally,
weight is a force, so for an object that weighs w lbs, its mass can be
determined from the equation w = mg.
Exercises
In answering
√ the following questions, do not numerically evaluate numbers
such as 2, π/3, e2 , ln(4/3), etc. The exception to this is when the
Exercises 73
Damping
10. It is often stated that “the key difference between critical damping and
overdamping is that critical damping provides the quickest approach
to zero amplitude.” However, this statement is not true. This problem
investigates this for the case of when m = 1, k = 4, u(0) = 1, and
u′ (0) = −4.
a) Find the solution when c = 5, which is the over-damped case, and
when c = 4, which is the critically damped case. Sketch both
solutions on the same axes. Explain why the statement is not true.
b) Solve the two problems in part (a) but use the general initial condi-
tions u(0) = u0 and u′ (0) = u′0 . Use this to explain how to modify
the statement so that it is true.
11. It is usually stated that negative damping is unstable. For the mass-
spring-dashpot system, negative damping means that c is negative.
From the solution, explain why the system is unstable for any nonzero
initial conditions.
Resonance and Forced Motion
12. A block weighing 4 lb stretches a spring 1.5 in. Assume that the block
is acted on by a periodic forcing as in (3.64), with F = 3 lb and ω =
16 /sec. At the start, the block is not moving and it is at its rest
position. Assume that g = 32 ft/s2 .
a) What IVP does u(t) satisfy?
b) What is the solution of the IVP?
c) Sketch the solution for 0 ≤ t ≤ π.
13. Suppose that a spring-mass system is at rest but, starting at t = 0,
the mass is subjected to a force of 5 cos 3t N. Assume that the mass is
2 kg, and the spring constant is 18 kg/s2 .
a) What IVP does u(t) satisfy?
b) What is the solution of the IVP?
c) Sketch the solution for 0 ≤ t ≤ 4π.
14. Suppose the forcing in (3.66) is replaced with F sin ωt. Does this
change (3.67)?
15. This exercise considers what happens when the forcing in (3.66) con-
sists of a combination of driving frequencies.
a) Suppose the forcing is
where the Fi ’s are nonzero, and the ωi ’s are all different, with ω0
given in (3.50). Does resonance still occur?
b) Suppose the forcing is F0 cos ω0 t cos ω1 t, where F0 is nonzero, ω1 6=
ω0 , and ω0 given in (3.50). Does resonance still occur?
76 Chapter 3. Second-Order Linear Equations
Complex Roots
In this case, the roots can be written as r = λ ± iµ, where
1
λ = (1 − b), (3.74)
2
and
1p
µ= 4c − (1 − b)2 . (3.75)
2
It is assumed here that 4c > (1 − b)2 . Writing the general solution as
in (3.73), and then separating into real and complex parts using Euler’s
formula, one finds that the resulting general solution can be written as
3.11.1 Examples
y = c1 x−3 + c2 x2 .
Example 2: Find the general solution of 4x2 y ′′ +17y = 0, for 0 < x < ∞.
Answer: Substituting in y = xr , one gets the equation 4r2 − 4r +
17 = 0. The solutions of this are r = 21 ± 2i. So, from (3.76), the
general solution is
√ √
y = d1 x cos(2 ln x) + d2 x sin(2 ln x).
Exercises
1. Assuming x > 0, find the general solution of the following Euler equa-
tions.
78 Chapter 3. Second-Order Linear Equations
2. Find the solution of the following problems. Before doing these prob-
lems, you might want to review Exercise 3, on page 63.
a) x2 y ′′ − 2xy ′ + 2y = x3 ex , where y(1) = 0, and y ′ (1) = 0
b) x2 y ′′ − 4xy ′ + 4y = −2x2 + 1, where y(1) = 0, and y ′ (1) = 0
c) x2 y ′′ − xy ′ + y = ln x, where y(1) = 0, and y ′ (1) = 0
d) xy ′′ + y ′ = x, where y(1) = 1, and y ′ (1) = −1
e) (x − 1)2 y ′′ + (x − 1)y ′ − y = 0, where y(2) = 1, and y ′ (2) = 0
Linear Systems
This chapter, and the one that follows, consider problems that involve
two or more first-order ordinary differential equations. Together the equa-
tions form what is called a first-order system. These are very common.
To explain why, it is worth considering a couple of examples.
Example 1: Mechanics
As stated on several occasions earlier in this text, one of the biggest
generators of differential equations is Newton’s second law, which states
that F = ma. To demonstrate its connection with a system of differential
equations, let x(t) denote the position of an object. The velocity is then
v = x′ (t), and the acceleration is a = x′′ (t). So, F = ma can be written
as mv ′ = F . Along with the equation x′ = v, the resulting system is
dx
= v,
dt
dv 1
= F.
dt m
As an example, for a uniform gravitation field, and including air resis-
tance, then F = −mg − cv (see Section 2.3.2). In this case, the system
becomes
x′ = v,
c
v ′ = −g − v.
m
This is a linear first-order system for x and v. It is also inhomogeneous
since x ≡ 0 and v ≡ 0 is not a solution.
Introduction to Differential Equations, M. H. Holmes, 2020
79
80 Chapter 4. Linear Systems
Example 2: Epidemics
Epidemics, such as the black death and cholera, have come and gone
throughout human history. Given the catastrophic nature of these events
there is a long history of scientific study trying to predict how and why
they occur. One of particular prominence is the Kermack-McKendrick
model for epidemics. This assumes the population can be separated into
three groups. One is the population S(t) of those susceptible to the
disease, another is the population I(t) that is ill, and the third is the
population R(t) of individuals that have recovered. A model that accounts
for the susceptible group getting sick, the subsequent increase in the ill
population, and the eventual increase in the recovered population is the
following set of equations [Holmes, 2019]
dS
= −k1 SI,
dt
dI
= −k2 I + k1 SI,
dt
dR
= k2 I.
dt
Given the three groups, and the letters used to designate them, this is an
example of what is known as a SIR model in mathematical epidemiology.
For us, this is an example of a nonlinear first-order system for S, I, and
R. The reason it is nonlinear is the SI term that appears in the first two
equations.
y ′′ + 2y ′ − 3y = 0 (4.5)
y ′ = v,
v ′ = 3y − 2v.
where r1 = −3 and !
1
a1 = .
−3
where r2 = 1 and !
1
a2 = .
1
d
x = Ax, (4.7)
dt
where A is an n × n matrix, and x is an n-vector, given, respectively, as
x1
a11 a12 ··· a1n
a21 x
a22 ··· a2n 2
A= . .. .. and x=
.. .
.. . ··· . .
an1 an2 ··· ann xn
x = c1 x 1 + c2 x 2
Exercises
1. Write the following as x′ = Ax, making sure to identify the entries in
x and A. If initial conditions are given, write them as x(0) = x0 .
a) u′ = u − v d) u′ = u − v
v ′ = 2u − 3v v ′ = 2u − 3v
b) 2u′ = −u u(0) = −1, v(0) = 0
3v ′ = u + v e) u′ = 2u − w
c) u′ = u − v + 2w v′ = u + v + w
v′ = u 3w′ = 2v + 6w
w′ = −u + 5v u(0) = −1, v(0) = 0, w(0) = 3
2. For the following: a) Write the equation in the form x′ = Ax. b) Find
the general solution of the second-order equation and then write it in
vector form as x = c1 x1 + c2 x2 , where x1 = a1 er1 t and x2 = a2 er2 t .
Make sure to identify a1 , a2 , r1 and r2 .
a) y ′′ + 2y ′ − 3y = 0 c) 3y ′′ + 4y ′ + 3y = 0
b) 4y ′′ + 3y ′ − y = 0 d) y ′′ + 4y ′ = 0
84 Chapter 4. Linear Systems
b) ! !
3 1 1
x′ = x, x=3 et
2 2 −2
c) ! ! !
′
2 0 1 2t 0
x = x, x= e + e−3t
0 −3 0 2
d)
1
! !
2 1 cos t
′
x = 1
x, x= et/2
−1 2
− sin t
d
x = Ax, for t > 0. (4.8)
dt
From (4.6), as well as Exercise 2 in the previous section, we have an idea
of what the general solution of this equation looks like. Namely, if we are
able to find n linearly independent solutions x1 (t), x2 (t), . . ., xn (t), then
the general solution can be written as
In the second step of the algorithm, when solving (4.15), we are inter-
ested in finding the vectors that can be used to form the general solution
of this equation. To say this more mathematically, we want to find lin-
early independent solutions. For those you might not remember what this
is, the definition is given next.
c1 a1 + c2 a2 + · · · + ck ak = 0, (4.16)
we get that
! ! !
2 1 1 0 2−r 1
A − rI = −r = .
1 2 0 1 1 2−r
−a + b = 0,
a − b = 0.
where !
1
a1 = . (4.19)
1
88 Chapter 4. Linear Systems
For the second eigenvalue r2 = 1, one finds that the eigenvectors have the
form a = aa2 , where a is an arbitrary nonzero constant and
!
1
a2 = .
−1
The above test applies irrespective of whether the eigenvalues are real or
complex valued. It also is not limited to a 2 × 2 matrix, and holds in the
general case of when the matrix is n × n.
where ! !
1 0
a1 = and a2 = . (4.21)
0 1
To check on independence, we are not able to use the Different Eigen-
values Test given above because a1 and a2 are eigenvectors for the same
eigenvalue. To use the definition, note that
! ! !
1 0 c1
c1 a1 + c2 a2 = c1 + c2 = .
0 1 c2
where !
−2i
a1 = .
1
Similarly, for r2 = 1 − i, one finds that the eigenvectors are
a = b a2 ,
where !
2i
a2 = .
1
Finally, because a1 and a2 are eigenvectors for different eigenvalues, they
are independent.
90 Chapter 4. Linear Systems
where !
1
a1 = .
0
Exercises
1. Determine whether the following pairs of vectors are linearly indepen-
dent.
! ! ! !
1 2 2 −1
a) a1 = , a2 = c) a1 = , a2 =
2 1 −8 4
! ! ! !
1 −3 −5 1
b) a1 = , a2 = d) a1 = , a2 =
−1 3 10 2
Example
The matrix in the differential equation,
!
1 2
x′ = x,
− 12 1
Because complex numbers are used for the r’s, both c1 and c2 must be
allowed to be complex-valued.
Given that x is real-valued, the coefficients c1 and c2 must be complex
conjugates. In other words, if c1 = α + iβ, where α and β are real-valued,
then it must be that c2 = α − iβ. We are going to separate the solution
into real and imaginary parts, which for the eigenvectors means that
! ! ! ! ! !
−2i 0 −2 2i 0 −2
= +i , and = −i .
1 1 0 1 1 0
where ! !
0 −2
p= , and q = .
1 0
Now, using Euler’s formula (3.14), we have that
General Formula
To summarize what was done in the above example, suppose that A is a
2 × 2 matrix with complex-valued eigenvalues r1 = λ + iµ and r2 = λ − iµ,
where λ and µ are real-valued with µ 6= 0. Also, assume that their
respective eigenvectors are a1 = p + iq and a2 = p − iq, where p and q
are vectors containing only real numbers. In this case, instead of writing
the general solution as
x = c1 a1 er1 t + c2 a2 er2 t ,
it can be written as
where
b1 = p cos(µt) − q sin(µt),
b2 = p sin(µt) + q cos(µt),
a solution of the form x = btert . However, this does not work, and to
find a second independent solution, the assumption is that
x = atert + bert .
where
b1 = p cos(µt) − q sin(µt),
b2 = p sin(µt) + q cos(µt).
(A − rI)b = a. (4.30)
4.5. Summary for Solving a Homogeneous Equation 95
where !
1
a1 = .
−1
(1 − i)a + b = 0
−2a − (1 + i)b = 0.
Both equations lead to the conclusion that b = −(1 − i)a. So, the
eigenvectors are
! !
a 1
a= =a = a a1 .
b −1 + i
x = d1 b 1 e t + d2 b 2 e t ,
where ! !
1 0
b1 = cos t − sin t,
−1 1
and ! !
1 0
b2 = sin t + cos t.
−1 1
Step 3: Satisfy the initial condition. Setting t = 0 in the general
solution, we get that
! ! !
1 0 0
d1 + d2 = .
−1 1 1
4.5. Summary for Solving a Homogeneous Equation 97
d1 = 0
−d1 + d2 = 1.
So, d1 = 0 and d2 = 1.
Step 4: The resulting solution is
" ! ! #
1 0
x(t) = sin t + cos t et .
−1 1
A solution of this is !
0
b= .
1
Step 2: Find the general solution. Using (4.29), the general solu-
tion is ! " ! !#
1 2t 1 0
x = c1 e + c2 t + e2t .
1 1 1
Step 3: Satisfy the initial condition. Setting t = 0 in the general
solution, we get that
! ! !
1 0 −1
c1 + c2 = .
1 1 2
Exercises
1. Find a general solution of the following differential equations.
! !
−1 6 0 −9
a) x′ = x f) x′ = x
1 0 1 0
! !
1
0 4 1 5
b) x′ = x g) x′ = x
1 0 − 41 −1
! !
1
2 1 1 4
c) x′ = x h) x′ = x
6 3 −5 0
!
2 0
!
′
d) x = x 1 3
−1 2 i) x′ = x
−1 3
! !
−2 0 0 0
e) x′ = x j) x′ = x
0 −2 0 0
2. Find the solution of the initial value problem x′ = Ax, where the
differential equationis given in the previous problem, and the initial
4
condition is x(0) = −1 .
3. A solution of x′ = Ax is given below. What are the eigenvalues of A,
and what are corresponding eigenvectors?
! ! !
1 3t 1 e −2t
a) x(t) = e +2 et c) x(t) =
1 −1 3e4t
! ! !
1 −5t 1 e−8t − e−t
b) x(t) = e + d) x(t) =
0 3 3e−t
4. The general solution (4.25), and the eigenvalue algorithm given in Sec-
tion 4.3, can be used for any dimension n. In this exercise you are to
find the general solution for the case of when n = 3.
0 1 1 −2 2 0
a) x′ = 1 0 1x c) x′ = 0 1 0 x
1 1 0 4 2 −1
1 1 2 −1 0 0
b) x′ = 0 2 0x d) x′ = 0 1 2 x
0 1 1 0 2 −1
4.6. Phase Plane 99
4.6.1 Examples
Two Positive Eigenvalues.
y-axis
0 0
-5 -5
-10 -5 0 5 10 -10 -5 0 5 10
x-axis x-axis
y-axis
0 0
-5 -5
-10 -5 0 5 10 -8 -4 0 4 8
x-axis x-axis
y-axis
0 0
-1 -1
-1 0 1 -1 0 1
x-axis x-axis
Table 4.1. Examples of integral curves and how they depend on the eigenval-
ues of A. Each curve corresponds to a specific choice for the constants appearing in
the general solution. The arrows indicate the direction for increasing t. It is assumed
here that µ 6= 0.
4.6. Phase Plane 101
1
c1 = 0 : Since x = c2 −1 et , then x = c2 et and y = −c2 et . In other
words, y = −x. This is the red line in Table 4.1(a) with negative
slope. Because et increases with t, the solution moves outward, away
from the origin. So, the arrows on the line point outward. Note that
1
the line is determined by the eigenvector a2 = −1 , and the direc-
tion on the line is determined by the positivity of the corresponding
eigenvalue r2 = 1.
3
c1 = 0 : Since x = c2 −2 e−3t , then x = 3c2 et and y = −2c2 et . In other
words, y = −2x/3. This is the red line in Table 4.1(b) with negative
slope. Because e−3t decreases with t, the solution moves inward,
toward from the origin. So, the arrows on the line point inward.
3
Note that the line is determined by the eigenvector a2 = −2 , and
the inward direction on the line is determined by the negativity of
the corresponding eigenvalue r2 = −3.
Imaginary Eigenvalues.
When the eigenvalues are imaginary, the integral curves are concentric
ellipses centered at the origin (see Exercise 5). To demonstrate this,
consider the differential equation
!
′ −2 4
x = x.
−2 2
4.6. Phase Plane 103
The eigenvalues are r1 = 2i and r2 = −2i, and the general solution, from
(4.28), is
" ! ! # " ! ! #
2 0 2 0
x(t) = d1 cos 2t− sin 2t +d2 sin 2t+ cos 2t . (4.34)
1 1 1 1
Complex Eigenvalues.
When the eigenvalues have nonzero real and imaginary parts the integral
curves are spirals centered at the origin (see Exercise 5). As an example,
!
′ 2 1
x = x,
−10 0
and these are responsible for the motion around the origin. This is similar
to what happens when r = ±iµ. However, these terms are multiplied by
eλt , and this causes the radial distance from the origin to either increase,
when λ > 0, or decrease, when λ < 0.
where !
6
x(0) = . (4.38)
−3
This is the same differential equation used for the phase plane example in
Table 4.1(b), and the general solution is given in (4.32). From the initial
condition, the solution is found to be
! !
3 1 2t 9 3
x(t) = e + e−3t . (4.39)
5 1 5 −2
The plot of this curve in the phase plane is shown in Figure 4.1. The
integral curves for the differential equation, which appear in Table 4.1, are
also included in the figure. As this shows, the solution of the initial value
problem is simply a portion of one of its integral curves. The starting
5
y-axis
-5
-10 -5 0 5 10
x-axis
Figure 4.1. The solid blue curve is the solution (4.39), and the solid blue dot
is the location of the initial condition (4.38). The dashed blue curves, and the red lines,
are integral curves for (4.37).
Exercises 105
Exercises
1. Phase portraits are shown in Figure 4.2, with arrows on some of the
curves. Do, or answer, the following: (i) Draw arrows on the other
curves. (ii) What properties of the eigenvalues result in the integral
curves shown in the phase portrait? (iii) Three different initial condi-
tions are shown by the black dots. For each one, sketch the solution
for the resulting IVP.
2. The eigenvalues for the following equations are real-valued. You are
to sketch the phase portrait as follows: (i) Draw the (red) lines that
are determined from the eigenvectors, and include the four arrows. (ii)
In each of the four quadrants determined by the red lines, include two
integral curves, with arrows.
! !
3 −1 3 2
a) x′ = x c) x′ = x
−1 3 −4 −3
! !
′
−6 3 ′
4 −2
b) x = x d) x = x
−4 1 3 −3
3. The eigenvalues for the following equations are imaginary. You are
to sketch the phase portrait as follows: draw three concentric ellipses
centered at the origin with arrows indicating the direction of motion.
It is useful to know that for A given in (4.4), when the eigenvalues
are imaginary, the elliptical integral curves are tilted right, as in Table
4.1(d), if ab < 0, and they are tilted left, as in Figure 4.2(e), if ab > 0.
! !
−1 2 2 1
a) x′ = x c) x′ = x
−2 1 −6 −2
! !
′
3 6 ′
−3 −1
b) x = x d) x = x
−3 −3 12 3
5 5
y-axis
y-axis
0 0
-5 -5
-10 -5 0 5 10 -10 -5 0 5 10
x-axis x-axis
5 5
y-axis
y-axis
0 0
-5 -5
-10 -5 0 5 10 -10 -5 0 5 10
x-axis x-axis
5 5
y-axis
y-axis
0 0
-5 -5
-10 -5 0 5 10 -10 -5 0 5 10
x-axis x-axis
Figure 4.2. Integral curves, and location of three initial conditions, for Exercise 1.
y-axis
y-axis
0
0 0
x-axis x-axis
Figure 4.3. Left-handed (on the left) and right-handed (on the right) spirals.
4.7 Stability
The phase plane can be useful for visualizing stability or instability of
a steady state solution. To explain how, recall from Section 2.4 that a
steady state is a constant that satisfies the differential equation. So, for
the equation x′ = Ax, a steady state is a constant vector xs that satisfies
Axs = 0. To avoid complications, it will be assumed that A is invertible,
which means that the only steady state solution is xs = 0. It is useful to
know that A is invertible if, and only if, r = 0 is not an eigenvalue for A.
The definitions of unstable and asymptotically stable are effectively
the same as in Section 2.4. Namely, a steady state xs is asymptotically
stable if any initial value x(0) chosen near xs results in
lim x(t) = xs . (4.40)
t→∞
The steady state is unstable if, no matter how close to xs you restrict
the choice for x(0), it is always possible to find an initial value x(0) that
results in the solution x(t) becoming unbounded as t increases.
It is easy to determine stability using the phase plane. For example,
in Table 4.1(a), when r1 > 0 and r2 > 0, the arrows on the integral
curves indicate movement out away from the origin. Consequently, this
is an example of when xs = 0 is unstable. Conversely, when r1 < 0 and
r2 < 0, the flow in towards the origin, and this means xs = 0 is asymp-
totically stable. In fact, looking at the various possibilities in Table 4.1,
you conclude that if A has an eigenvalue with Re(r) > 0, then the steady
state in unstable. Similarly, if the eigenvalues of A are both negative, or
if Re(r) < 0, then the steady state is asymptotically stable.
108 Chapter 4. Linear Systems
The conclusions in the previous paragraph were made using the phase
portraits in Table 4.1. For those that prefer more rigorous derivations,
then the formulas for the general solutions given in Section 4.5 can be
used.
Our classification of a steady state being unstable or asymptotically
stable does not include what happens when the eigenvalues are imaginary.
As shown in Table 4.1(d), the solution does not decay to zero, or blowup,
but simply encircles the origin. In this case, the steady state is said to be
neutrally stable.
The other case we are missing here is what happens when the matrix is
defective. From (4.29), the conclusion we had earlier still holds. Namely,
if r < 0, then we have asymptotically stability, and if r > 0, then we have
instability.
The above discussion is summarized in the following theorem.
Stability Theorem for a Linear System. For x′ = Ax, if r = 0 is
not an eigenvalue for A, then the following hold:
1. If all of the eigenvalues of A satisfy Re(r) < 0, then xs = 0 is an
asymptotically stable steady state.
2. If A has one or more eigenvalues with Re(r) > 0, then xs = 0 is
an unstable steady state.
3. If the eigenvalues of A are imaginary, then xs = 0 is a neutrally
stable steady state.
It is worth pointing out that the first two conclusions in the above theorem
hold when A is n × n. The third one also holds for the n × n case if
you make the additional assumption that A is not defective. For those
who might be wondering what happens when r = 0 is an eigenvalue,
the solution of Ax = 0 is no longer just xs = 0. In fact, any and all
eigenvectors for r = 0 are steady state solutions. It is possible to examine
the various cases that arise in this situation related to stability, but this
will not be considered in this text.
In addition to their stability, steady states are often identified by the
geometric properties of the solution near the steady state. So, for example,
because of the outward direction of the flow in Figure 4.1(a), the steady
state is called a source. In contrast, because of the inward flow in Figure
4.1(c), the steady state is called a sink. For similar reasons, the flow in
Figure 4.1(e) is a spiral source, and the one in Figure 4.1(f) is a spiral
sink. Finally, the steady state in Figure 4.1(b) is a saddle, and the one
in Figure 4.1(d) is a center.
Example 3: Find the steady state, and determine its stability for
! !
′ 1 1 2
u = u+ . (4.41)
1 −1 4
Exercises
1. Determine whether xs = 0 is an asymptotically stable, unstable, or
neutrally stable steady state for the following differential equations.
Also, state whether the steady state is a sink, source, spiral sink, spiral
source, saddle, or center.
! ! !
−1 6 2 1 2 5
a) x′ = x d) x′ = x g) x′ = x
1 0 3 4 −5 −6
! ! !
1 2 −1 −1 0 −9
b) x′ = x e) x′ = x h) x′ = x
−3 −4 6 −6 1 0
! ! !
′
3 1 ′
1 14 ′
1 −4
c) x = x f) x = x i) x = x
1 3 −5 0 1 −1
Nonlinear Systems
Example 1: Pendulum
The equation for the angular deflection of a pendulum is (see Figure 5.1)
d2 θ
ℓ = −g sin θ, (5.1)
dt2
where the initial angle θ(0) and the initial angular velocity θ′ (0) are as-
sumed to be given. Also, ℓ is the length of the pendulum and g is the grav-
itational acceleration constant. Introducing the angular velocity v = θ′
then the equation can be written as the first-order system
θ′ = v, (5.2)
′
v = −α sin θ, (5.3)
111
112 Chapter 5. Nonlinear Systems
Example 2: Measles
A model for the spread of a disease, like measles, is
dS
= αN − (βI + α)S,
dt
dI
= βIS − (α + γ)I.
dt
The equations for the pendulum and the spread of measles are not
solvable using elementary functions. What is possible it to ask questions
about the solution that are significant and answerable. As an example,
with measles, a reasonable question would be: what would it take to
eliminate the disease from the population? This requires that I → 0 as
t → ∞ (and the faster this happens the better). In more mathematical
terms, we want I = 0 to be an asymptotically stable steady state. How
to modify the stability of I = 0, with the goal of quickly eliminating the
disease, will be considered in Section 5.2.2.
A question arising with the pendulum is, does it ever stop moving?
Given the physical assumptions used in the derivation of the equation it
is reasonable to expect that it does not stop and, in fact, the solution
is expected to be periodic. So, we would like to know if it is possible
to show that the solution is periodic, and in the process determine the
period (without actually solving the problem).
In these equations, u(t) and v(t) are the dependent variables, and f and
g are given functions of u and v. It is assumed that the equations are
autonomous, which means that f and g do not depend explicitly on t.
The vector form of (5.4), (5.5) is
dy
= f (y), (5.6)
dt
where ! !
u f (u, v)
y= , and f= .
v g(u, v)
For an initial value problem, an initial condition of form
!
u0
y(0) = . (5.7)
v0
would also be given.
5
v-axis
-5
-3 -2 -1 0 1 2 3
u-axis
Figure 5.2. Solution curves of (5.8), (5.8) in the u,v-plane for four different
initial conditions (shown with the solid dots). The arrows indicate the direction for
increasing t.
114 Chapter 5. Nonlinear Systems
f (us , vs ) = 0, (5.10)
g(us , vs ) = 0. (5.11)
3 − u − v − uv = 0,
uv − 2v = 0.
1 − (1 + α)u + u2 v = 0,
u − u2 v = 0.
Exercises 115
x′ = x − x2 − xy,
y ′ = 2y − y 2 − 3xy.
x − x2 − xy = 0,
2y − y 2 − 3xy = 0.
x′ = x − y,
y ′ = (x − y)3 ,
Exercises
1. Write the following as y′ = f (y), making sure to identify the entries in
y and f . If initial conditions are given, write them as y(0) = y0 .
116 Chapter 5. Nonlinear Systems
a) u′ = u2 − v g) Michaelis-Menten system
v ′ = 2u − 3v S ′ = −k1 ES + k−1 (E0 − E),
b) u′ = u2 + v 2 E ′ = −k1 ES+(k2 +k−1 )(E0 −E)
2v ′ = sin(u) S(0) = 1, E(0) = 2
c) u′ = eu − v h) Predator-prey
v ′ = uv x′ = ax − bxy
u(0) = −1, v(0) = 0 y ′ = −cy + dxy
d) Van der Pol oscillator i) Projectile (nonuniform field)
gR2
u′′ + (1 − u2 )u′ + u = 0 y ′′ = −
(R + y)2
e) Toda oscillator y(0) = 0, y ′ (0) = 3
u′′ + eu − 1 = 0
j) Orbital motion
f) Duffing oscillator α2 µ
u′′ + u + u3 = 0 r′′ = 3 − 2
r r
u(0) = 1, u′ (0) = −1 r(0) = 1, r′ (0) = 2
5.2 Stability
The question considered now is central to this chapter, and it is whether
a steady state is achievable. What this means is that the steady state is
asymptotically stable. To explain how we are going to determine stability,
consider the problem of solving
x′ = x − x2 − xy, (5.12)
′ 2
y = 2y − y − 3xy. (5.13)
This is the problem from Example 3 in the previous section, and we found
that there are four steady states: (0, 0), (0, 2), (1, 0), and (1/2, 1/2). One
approach for providing insight about stability is to solve the problem
numerically. This is easy to do, and two computed solution curves are
shown in Figure 5.3. The curves are consistent with what is expected if
5.2. Stability 117
2
y-axis
0
-2
-4
-3 -2 -1 0 1 2 3 4
x-axis
(0, 2) and (1, 0) are asymptotically stable. Also, since both curves start
near (0, 0), yet move away from it, it would not be a surprise to find out
that (0, 0) is an unstable steady state.
Solving the problem numerically is so easy that it possible to solve the
problem for many different initial conditions, and check if the solution
approaches one of the various steady states. The results from such a
calculation are shown in Figure 5.4. What is found is that there are,
apparently, two asymptotically stable steady states, (0, 2) and (1, 0). The
calculations also identity the regions for the initial conditions that result
in the solution ending up at the respective steady state. The two regions
determined from this computation are called the domain of attraction for
the respective steady state.
Our goal is not to be able to determine the shaded regions shown
in Figure 5.4, but, rather, to show that there is a small region around
the respective steady state with the same property as the shaded region.
Namely, for any initial condition in that small region, the solution of the
resulting IVP will end up at the steady state. In this case, the steady
state is said to be asymptotically stable. What we are doing now is the
two dimensional version of what we did in Section 2.4, and the nonlinear
version of what was done in Section 4.7.
Figure 5.4. An initial condition (x(0), y(0)) located in one of the shaded
regions results in the solution of (5.12),(5.13) ending up at the steady state in that
shaded region. The two steady states are shown by the dark circles.
f (us , vs ) = 0,
g(us , vs ) = 0.
The reason for considering stability comes from this question: If we start
the solution near (us , vs ), what happens?
There are three possible conclusions coming from this question: the
steady state is unstable, it is asymptotically stable, or it is neutrally
stable. What these are can be explained using a ball and bowl (see Figure
5.5). The force on the ball is gravity. For the bowl, the steady state is at
the bottom, and for the inverted bowl it is at the top. For the inverted
bowl, if you release the ball from rest, no matter where you place it (other
than exactly at the top), the ball will roll away. The conclusion is that
the steady state is unstable. For the bowl, you can control how far the
ball will get from the steady state (the bottom) by placing it close to
the bottom and giving it only a small initial velocity. Consequently, the
steady state is stable. Because the only force is gravity, the ball will roll
around in the bowl forever. This means the steady state is neutrally
stable. If the problem also includes damping, such as friction, then the
ball will slow down and eventually come to rest at the bottom. In this
case the steady state is asymptotically stable. Note that including
Figure 5.5. Ball in a bowl, on the left, and a ball on an inverted bowl, on the right.
5.2. Stability 119
damping for the inverted bowl will not change the fact that the top is an
unstable steady state.
For those that prefer a more mathematical definition, the idea under-
lying asymptotic stability is that if y(0) is any point close to the steady
state ys , then
lim y(t) = ys . (5.16)
t→∞
As stated above, a steady state is stable if you can control how far the
solution gets from ys by picking y(0) close to ys . Specifically, given any
ε > 0, you can find a δ > 0 so that if ||y(0)−ys || < δ, then ||y(t)−ys || < ε.
If this is not possible then ys is unstable. If ys is stable, and (5.16) holds,
then it is asymptotically stable. Otherwise it is said to be neutrally stable.
Note that this version of the definition of stability√requires that u and v
have the same physical dimensions so that ||y|| = u2 + v 2 is defined.
To answer the stability question, assume that the initial position
(u(0), v(0)) is very close to (us , vs ). To determine what happens, we will
use what is called the linear approximation in multivariable calculus. This
states that if f (u, v) and g(u, v) are differentiable at (us , vs ), then each
can approximated using their respective tangent plane. In particular,
y′ = J(y − ys ), (5.17)
where ! !
u us
y= , ys = ,
v vs
and !
fu (us , vs ) fv (us , vs )
Js = .
gu (us , vs ) gv (us , vs )
The matrix Js is known as the Jacobian matrix of f evaluated at ys .
120 Chapter 5. Nonlinear Systems
To put the problem into the form covered in the last chapter, let
x = y − ys . With this, (5.17) becomes
x′ = Ax, (5.18)
There is a notable hole in the above list in that there is no conclusion for
the case of when the eigenvalues are imaginary. In the theorem in Section
4.7, this is referred to as being neutrally stable. There are neutrally stable
steady states for nonlinear systems, as illustrated with the ball and bowl
example earlier, but the tangent plane approximation is inadequate to
determine this. One approach to show neutral stability is to use the ideas
developed in Section 5.3.
As a final comment, the only assumption needed to guarantee that the
above conclusions hold is that the first and second partial derivatives of
f (u, v) and g(u, v) are continuous. Those interested in a mathematically
rigorous proof of this should consult Stuart and Humphries [1998] or Perko
[2001].
Phase Plane
The above derivation for the stability conditions can provide us with in-
formation about the solution curves near a steady state. The reason is
that the reduced equation in (5.18) is the same one considered in the last
5.2. Stability 121
chapter. This enables us, in certain cases, to apply the phase plane solu-
tions shown in Table 4.1 (page 100) to the nonlinear system. To explain
how, suppose you have a steady state that the above test determines is
unstable or asymptotically stable. As stated earlier, we are only consid-
ering isolated steady states, and to guarantee this happens it is assumed
that r = 0 is not an eigenvalue. Now, in the vicinity of the steady state,
we have that y ≈ ys +x. This means that the phase portrait for y is simi-
lar to one of those in Table 4.1, but it is centered at y = ys rather than at
x = 0. Which one is determined by the eigenvalues of Js . Demonstrations
of this will be included in the examples that follow.
5.2.2 Summary
For the nonlinear system
u′ = f (u, v)
v ′ = g(u, v),
5.2.3 Examples
√ √
and this has eigenvalues r1 = (−1 + i 29)/2 and r2 = (−1 − i 29)/2.
Since both satisfy Re(r) < 0, this steady state is asymptotically stable.
In addition, since the eigenvalues are complex, and Re(r) < 0, the
phase portrait near this steady state should be a spiral sink. To check,
the region in Figure 5.2 that is near (2α, α) is shown in Figure 5.6.
As expected, the solution curves spiral into the steady state, as they
should for a spiral sink.
(0, 0): In this case !
− 21 1
Js = ,
4 − 12
and this has eigenvalues r1 = 3/2 and r2 = −5/2 . Since r1 > 0 then
this steady state is unstable. Also, since r2 < 0 < r1 , then this is a
saddle point and the phase portrait near (0, 0) will resemble the one in
Figure 4.1(b) or in Figure 4.2(b).
−(2α, α): In this case !
− 12 1
Js = ,
− 29
4 − 12
√ √
and this has eigenvalues r1 = (−1 + i 29)/2 and r2 = (−1 + i 29)/2.
Since both satisfy Re(r) < 0, this steady state is asymptotically stable.
As with (2α, α), this is a spiral sink.
x′ = x − x2 − xy,
y ′ = 2y − y 2 − 3xy.
This is the system that produced the solution curves shown in Figure
5.3.
Answer: In Section 5.1.1, Example 3, we found that there are four
2
v-axis
-0.3
0.6 1 1.4
u-axis
Figure 5.6. Solution curves of (5.8), (5.8) in the u,v-plane near the steady
state (2α, α).
124 Chapter 5. Nonlinear Systems
steady states: (0, 0), (0, 2), (1/2, 1/2) and (1, 0). To determine their
stability, the Jacobian is
∂f ∂f !
∂x ∂y 1 − 2x − y −x
J= ∂g
= .
∂g −3y 2 − 2y − 3x
∂x ∂y
The two red lines shown in Figure 5.7 are determined by these eigen-
vectors. The arrows point toward the steady state as both eigenvalues
are negative. Typical integral curves are shown in blue. The result is
a phase portrait for a sink.
(1/2, 1/2): In this case
!
−1/2 −1/2
Js = ,
−3/2 −1/2
√ √
and this has eigenvalues r1 = (−1+ 3)/2 and r2 = (−1− 3)/2. Since
r1 > 0, it follows that this steady state is unstable. As for the phase
portrait near this steady state, since r2 < 0 < r1 , then this steady
2.1
y-axis
1.9
-0.02 0 0.02
x-axis
Figure 5.7. Phase portrait near the steady state (0, 2) for Example 2.
5.2. Stability 125
0.52
y-axis 0.5
0.48
0.48 0.5 0.52
x-axis
Figure 5.8. Solution curves of Example 2 in the x,y-plane near the steady
state (1/2, 1/2).
The two red lines determined by these vectors are shown in Figure
5.8. Typical integral curves are shown in blue. So, the curves have the
pattern expected for a saddle.
Determining the stability of the remaining two steady states is left as
an exercise.
(S, I) = (Se , Ie ): One finds that this steady state is unstable if N < Se ,
and it is asymptotically stable if N > Se .
Exercises
1. For the following find the steady states, and then determine whether
they are asymptotically stable, unstable, or indeterminate. Also, ex-
cept for the indeterminate cases, state whether the steady state is a
sink, source, spiral sink, spiral source, or saddle. Any parameters ap-
pearing in the equations should be assumed to be positive.
Exercises 127
( (
u′ = 1 − 2u − v − uv x′ = ex − y
a) g)
v ′ = 3uv − v y ′ = xy
(
( x′ = ax − bxy
u′ = v − u h)
b) y ′ = −cy + dxy
v ′ = v + u3
(
2SP
(
u′ =1+v S ′ = 2S − S 2 − 1+S
c) i) 2SP
v′ = u + v3 P′ = 1+S −P
(
S ′ = − 21 IS + 1 − I − S
(
u′ =4− uv 2 j)
d) I ′ = 12 IS − I
v ′ = −v + uv 2
(
r′ = s − r
(
x′ = x2 − y
e) k)
y ′ = 2x − 3y s′ = (2 − r − s)(1 + s2 )
( (
x′ = x2 + y 2 S ′ = −2ES + E0 − E
f) l)
2y ′ = sin(x) E ′ = −2ES + 2(E0 − E)
2. For the following: (i) find the steady state, (ii) find the linear approx-
imation of the system near the steady state, and then (iii) sketch the
phase portrait in the vicinity of the steady state as follows: draw the
(red) lines that are determined from the eigenvectors of Js , including
the arrows for these lines, then in each of the four quadrants deter-
mined by the red lines, include two integrals curves, with arrows.
( (
u′ = v − u r′ = s − r
a) c)
v ′ = v + u3 s′ = (2 − r − s)(1 + s2 )
( (
u′ = 1 + v S ′ = −2ES + E0 − E
b) d)
v′ = u + v3 E ′ = −2ES + 2(E0 − E)
The trace of a matrix is the sum of the numbers on the diagonal. The
formula is tr(Js ) = a + d. Also, the determinant is det(Js ) = ad − bc.
128 Chapter 5. Nonlinear Systems
h i
1
p
a) Show that the eigenvalues of Js are 2 tr(Js )± [tr(Js )]2 − 4det(Js ) .
b) Show that if tr(Js ) > 0, then ys is unstable.
c) Show that if det(Js ) < 0, then ys is unstable.
d) Show that if det(Js ) > 0 and tr(Js ) < 0, then ys is asymptotically
stable.
5. This exercise considers the curve, in the first quadrant, that separates
the red and blue regions in Figure 5.4.
a) Explain why the curve must contain the point (1/2, 1/2).
b) Suppose that the initial point (x(0), y(0)) is on the curve. Explain
why the resulting solution (x(t), y(t)) must remain on the curve.
6. A model for how a joke moves through a population involves three
groups: S is the population that either has not heard the joke, or does
not remember it, T is the population of those who know the joke and
they will tell it to others, and R is the population who know the joke
but will not tell it to others (they are not good joke tellers or they
don’t think it’s all that funny). As shown in Holmes [2019],
dS
= −2αST + β(N − S),
dt
dT
= αST − βT,
dt
where N is the total number of individuals in the population (it is
constant). The coefficients α and β are positive constants. Also, once
S and T are determined, then R = N − T − S.
a) There are two steady states, what are they?
b) One of the steady states has T = 0. When is it asymptotically
stable?
c) One of the steady states has T 6= 0. When is it asymptotically
stable?
d) The α is the telling parameter, so a larger α means the joke is
being told more often. Similarly, β is the forgetting parameter, so
a larger β means the joke is being forgotten faster. Based on your
answers from parts (b) and (c), under what conditions will the joke
disappear from the population?
in our lives, and examples are the sleep-wake cycle and the periodic events
associated with the Earth’s rotation.
To begin, it’s best to define what is meant by periodicity. A solution
of y′ = f (y) is periodic if there is a positive number T so that
y(t + T ) = y(t), ∀t ≥ 0. (5.22)
The smallest positive T , if it exists, is the period.
We will only consider problems that come from Newton’s second law.
Specifically, if u(t) is the displacement, and F is a function of u, then
F = ma gives us the differential equation
mu′′ = F (u). (5.23)
Letting v = u′ , then the above equation can be written in system form as
u′ = v, (5.24)
1
v ′ = F (u). (5.25)
m
It is not hard to show that if u(t) is periodic with period T , then the
velocity v(t) is also periodic with period T . Consequently, (5.22) is sat-
isfied, and so the solution is periodic. Examples of what are, or are not,
periodic are explored in more depth in Exercise 2.
We will first find a way to determine the solution curve in the phase
plane directly from the differential equation and initial conditions. Once
that is done, we will then be able to determine the period T , as well as
other properties of the solution.
Example: Mass-Spring
In Section 3.10, it was shown that the displacement u(t) of a mass in
a spring-mass system satisfies mu′′ + ku = 0. The general solution of
this equation can be written as p u = R cos(ω0 t − ϕ), and v = u′ =
−ω0 R sin(ω0 t − ϕ), where ω0 = k/m. Consequently, the solution is
periodic, with period T = 2π/ω0 . The key observation here is that, using
the identity cos2 θ + sin2 θ = 1,
u 2 v 2
+ = 1,
R ω0 R
or equivalently
1 2
u2 + v = R2 . (5.26)
ω02
This is an equation for an ellipse in the u,v-plane. As an example, suppose
that m = 1, k = 4, and the initial conditions are u(0) = 1 and v(0) = 0.
In this case, u = cos(2t), v = −2 sin(2t), and from (5.26), the ellipse is
1
u2 + v 2 = 1. (5.27)
4
130 Chapter 5. Nonlinear Systems
v-axis
0
-1
-2
-2 -1 0 1 2
u-axis
Figure 5.9. Elliptical path, given in (5.27), that is followed by the solution
of the mass-spring IVP. The blue dot is the location of the initial condition.
u′ = v,
v ′ = −ω02 u.
This can be used to determine the direction of the arrows in Figure 5.9.
Since v ′ = −ω02 u, using the initial condition given earlier, v ′ (0) = −ω02 .
The fact that this is negative means that v must decrease as it leaves
the initial point, and so the direction of motion is clockwise around the
curve. Note that it is not possible for the solution to reverse direction on
the curve because this would require that there is a point on the curve
where u′ = 0 and v ′ = 0. Such a point corresponds to a steady state, and
the only steady state for this problem is the origin.
mu′′ + ku = 0,
mvv ′ + kuu′ = 0.
It was mentioned earlier that not every forcing function will result in
a closed curve. For the above mass-spring problem the spring force is
F = −km. This is attractive, in the sense that it pulls the mass back
towards the rest position u = 0. If the force is repelling, so F = km,
then instead of (5.32), you get u2 − 41 v 2 = 1. This is an equation for a
hyperbola, which is not a close curve.
Example: Pendulum
The equation for the angular deflection of a pendulum can be written as
d2 θ
= −α sin θ. (5.33)
dt2
where α = g/ℓ. Introducing the angular velocity v = θ′ , then we obtain
the first-order system
θ′ = v, (5.34)
′
v = −α sin θ. (5.35)
5.3. Periodic Solutions 133
1.6
v-axis
0
-1.6
- /2 - /4 0 /4 /2
-axis
In this example, assume that α = 4, and that the initial conditions are
θ(0) = π/4 and v(0) = 0. To determine the closed solution curve, we
multiply (5.33) by the velocity v = θ′ , giving us
vv ′ = −4θ′ sin θ.
Writing this as
d 1 2 d
v = 4 cos θ ,
dt 2 dt
and then integrating gives us the equation
1 2
v − 4 cos θ = c.
2
√
With the initial conditions we find that c = −2 2, and so the equation
for the curve takes the form
√
v 2 − 8 cos θ = −4 2. (5.36)
v
M
/4
0
v
- /4
-v
M
0 2 4 6 8 10
t-axis
Figure 5.11. Solution curves for θ(t) and v(t) for the pendulum solution
shown in Figure 5.10.
0
v
-6
-3 -2 - 0 2 3
Figure 5.12. Phase portrait for the pendulum equations (5.34), (5.35), when α = 4.
5.3. Periodic Solutions 135
Example: Mass-Spring
The equation
√ for the curve is given in (5.32). Solving this for v yields
v = ±2 1 − u2 . Which sign you use depends on what part of the curve
you are considering. In Figure 5.9 the two u intercepts are u = ±1. So, for
the lower part
√ of the curve connecting (1, 0) to (−1, 0), v is negative, and
so v = −2 1 − u2 . Since v = u′ , then we have the first-order differential
equation
du p
= −2 1 − u2 .
dt
This equation is separable, which yields
du
Z Z
√ = −2dt.
1 − u2
Carrying out the integrations,
arcsin(u) = −2t + c.
Given that u = 1 at t = 0, then c = π/2.
To determine the period, we solve the above equation for t to obtain
1π
t= − arcsin(u) .
2 2
It is now possible to determine how long it takes for the solution to move
along the lower half of the curve and arrive at (−1, 0). Namely, letting
u = −1 in the above equation we get that
1π π
t= − arcsin(−1) = .
2 2 2
To compute the time to transverse the upper part of the curve, you can
either use the separation of variables approach or you can use the sym-
metry of the solution curve. Both yield the result that the time is π/2.
Therefore, the period is the sum, which means that T = π. This agrees
with what we found earlier using the exact solution to the problem.
136 Chapter 5. Nonlinear Systems
Example: Pendulum
The equation for the curve is given in (5.36). Solving this for v yields
p √
v = ±2 2 cos θ − 2. The lower part of the solution curve, shown in
Figure p
5.10, goes from (π/4, 0) to (−π/4, 0). On this part of the curve
√
v = −2 2 cos θ − 2, which gives us the first-order differential equation
√
q
dθ
= −2 2 cos θ − 2 .
dt
This equation is separable, which yields
dθ
Z
p √ = −2t + c.
2 cos θ − 2
In anticipation of imposing the initial condition, the above integral is
written as Z θ
dr
p √ = −2t + c.
θ0 2 cos r − 2
Now, given that θ(0) = π/4, then θ0 = π/4 and c = 0. The above
equation then takes the form
Z θ
dr
p √ = −2t.
π/4 2 cos r − 2
The time to reach θ = −π/4 is therefore
1 −π/4 dr
Z
t=− p √
2 π/4 2 cos r − 2
1 π/4 dr
Z
= p √ .
2 −π/4 2 cos r − 2
Exercises
1. Find a Hamiltonian function H(u, v) for each of the following:
v-axis
0
-1 1
u-axis
a) The path followed by the solution is shown in Figure 5.14. Find the
equation for this closed curve.
b) Find the steady-state, show that it is not on the curve you found
in part (a).
c) Draw arrows on the curve indicating the direction of motion. Make
sure to explain how you determine this.
d) What is the maximum velocity?
e) What is the minimum displacement?
f) Find a formula, similar to the one in (5.37), for the period.
5. The problem concerns what is known as a Morse oscillator, and the
differential equation is
Assume the initial conditions are u(0) = 1 and v(0) = 0. This equation
arises when studying the vibrational energy of a diatomic molecule.
a) The path followed by the solution is shown in Figure 5.15. Find the
equation for this closed curve.
b) Find the steady-state, show that it is not on the curve you found
in part (a).
c) Draw arrows on the curve indicating the direction of motion. Make
sure to explain how you determine this.
v-axis
0
u-axis
Figure 5.14. Path followed by the solution of the Duffing oscillator in Exercise 4.
5.4. Motion in a Central Force Field 139
v-axis
0
0
u-axis
Figure 5.15. Path followed by the solution of the Morse oscillator in Exercise 5.
Figure 5.16. A particle, the red dot, orbits a particle located at the origin.
The orbit curve lies in a plane containing the origin and has normal n, where n is
parallel to p = x0 × v0 .
mp2
mr′′ = f (r) + . (5.43)
r3
This is a force balance equation, where f (r) is the force introduced earlier
and mp2 /r3 is an outward directed force due to angular momentum.
5.4. Motion in a Central Force Field 141
r′ = v, (5.44)
1 p2
v′ = f (r) + 3 . (5.45)
m r
It is worth knowing that in the derivation of (5.43), it is found that
p = ||x0 × v0 ||. So, p is a positive constant that is known from the initial
conditions.
From this one finds that the eigenvalues are ±ip/rs2 , which means that the
stability of the steady state is indeterminate using the Linearized Stability
Theorem.
1 mp2 k
mv 2 + − = c, (5.46)
2 2r2 r
where c is a constant determined by the initial conditions. Completing
the square, we get that
1 1 2
v 2 + p2 − = c20 , (5.47)
r rs
142 Chapter 5. Nonlinear Systems
v-axis
v-axis
0 0
u- 0 u 0 u- u
+ +
u-axis u-axis
Figure 5.17. Two possible elliptical curves coming from (5.48). The u-
intercepts for each ellipse are u− and u+ .
2
where c20 = v02 + p2 1/r0 − 1/rs , r(0) = r0 , and v(0) = v0 .
To answer the question about a periodic orbit, it will make things
easier if we let u = 1/r. So, (5.47) takes the form
2
v 2 + p2 u − us = c20 , (5.48)
1.5
0.7
v-axis
y-axis
0
0
-1.5 -0.6
0 1 2 0 1 2
r-axis x-axis
Figure 5.18. Numerical solution of (5.43) in the case of when the solution is
periodic. The initial position, and direction of motion, are shown on each curve. The
two time points used to place the direction arrows on the left are the same time points
used on the right.
Exercises 143
on the solid curve, and no matter which direction you go on the curve, u
approaches zero. In other words, r → ∞. Physically, what is happening
is that the angular momentum is so large that an orbit is not possible,
and the particle simply escapes whatever hold the particle at the origin
might have on it. It is also evident from Figure 5.17, contrary to what
is often shown in cartoons, that the particle does not make several orbits
around the origin before escaping. In fact, the particle is incapable of
making even one complete orbit.
Exercises
1. Suppose the law of gravity results in f (r) = −k/r3 , where k > 0. You
can assume that k 6= mp2 .
a) Are there any steady state solutions? If so, check on their stability.
b) Assuming there is a periodic solution, determine its equation in the
u,v-plane.
c) Use your result from part (b) to explain why there is no periodic
solution of this problem.
2. Suppose the law of gravity results in f (r) = −kr, where k > 0. Note
that this is assuming that gravity acts like an elastic spring.
a) Are there any steady state solutions? If so, check on their stability.
b) Assuming there is a periodic solution, determine its equation in the
r,v-plane.
c) The solution curve is shown in Figure 5.19 in the case of when
r(0) = r0 and v(0) = 0. Show that the second r intercept is at
r0 (rs /r0 )2 , where rs is the steady state you found in part (a).
d) Where is the steady state located in Figure 5.19?
3. This problem concerns the solution shown in Figure 5.18.
a) In Figure 5.18(left), where is rs located?
b) In Figure 5.18(right), sketch in the circular orbit derived in Section
5.4.1.
v-axis
| |
0
r-axis
Laplace Transform
6.1 Definition
The generalization we are interested in called the Laplace transform, and
its definition is given next.
It will be useful to have a more compact notation for the integral in this
expression, and this will be done by writing the above formula as
145
146 Chapter 6. Laplace Transform
6.1.1 Requirements
There are a couple of ways to find the Laplace transform of a function.
One is to carry out the integration in (6.1), just as you did in calculus.
Much of the material in this section concerns the mathematical require-
ments needed to do this. However, it is also possible to find a Laplace
transform by looking it up in a table, such as the one in Table 6.1 (on
page 151). As demonstrated in Example 3 (version 2), this is very easy
to do. Most, but not all, of the transforms in this chapter can be done
using the given table (along with the formula in Exercise 4).
For the improper integral in (6.1) to exist, a condition must be im-
posed on the complex variable s. To explain, if y(t) = e3t , then using the
definition of an improper integral and (6.1)
Z T Z T
3t −st
Y (s) = lim e e dt = lim e(3−s)t dt (6.3)
T →∞ 0 T →∞ 0
1 (3−s)T 1
= lim e − . (6.4)
T →∞ 3−s 3−s
Clearly, we need s 6= 3. As for the limit, it is useful to know that, given
a nonzero complex number z,
(
zT 0 if Re(z) < 0,
lim e = (6.5)
T →∞ does not exist if Re(z) ≥ 0.
The proof of this comes directly from Euler’s formula, as expressed in
(3.15). For (6.4), z = 3 − s and this means that for the limit to exist we
need Re(3 − s) < 0, or equivalently, we need Re(s) > 3. In this case,
1
Y (s) = .
s−3
The requirement that Re(s) > 3 gives rise to what is known as the half-
plane of convergence for the Laplace transform.
The second mathematical requirement concerns the smoothness of
y(t). For the problems considered in this textbook, it is enough to as-
sume that y(t) is continuous for 0 ≤ t < ∞, except possibly for jump
discontinuities. What a jump discontinuity means is that y(t) is not
continuous at the point, but the limits of y from the left and right are
defined and finite (the two limits do not need to be equal). A simple
example, with a jump discontinuity at t = 2, is
(
3 if 0 ≤ t ≤ 2,
y(t) = (6.6)
−1 if 2 < t.
6.1. Definition 147
The specific requirement for the Laplace transform is that over any inter-
val 0 ≤ t ≤ T , y(t) is continuous except for possibly a finite number of
jump discontinuities. In this case, y(t) is said to be piecewise continu-
ous for t ≥ 0.
The final requirement on y(t) is to guarantee that the improper inte-
gral in (6.1) converges. Specifically, it is required that there is a constant
α so that
lim y(t)eαt = 0.
t→∞
If this holds, then y(t) is said to have exponential order. As examples, any
polynomial function in t, any linear combination of sin(ωt) and cos(ωt),
and any linear combination of terms of the form eωt have exponential
2 3
order. On the other hand, et and et do not. Throughout this chapter,
whenever taking the Laplace transform, it is assumed that the function
has exponential order.
In the above example it is stated that Re(s) > 3. This can not be
written as s > 3. The reason is that the usual definition of inequality can
not be used with complex numbers. For example, the rules of inequality
require that s2 ≥ 0. So, taking s = i, you end up concluding that −1 ≥ 0.
In fact, it is possible to prove that the complex numbers cannot be made
an ordered field no matter how you define the rule used for inequality.
6.1.2 Examples
With the technical details out of the way, we consider a few examples. As
you will see, finding the Laplace transform of a function provides ample
opportunity to practice using integration by parts. Also, in what follows
the improper integral will be treated as a definite integral, with the upper
endpoint being t = ∞ (see Example 1 below). It is understood that the
evaluation at the upper endpoint involves a limit, as expressed in (6.3).
1 s ∞
Z
∞
−st
= − cos(2t)e − cos(2t)e−st dt
2 t=0 2 0
1 s ∞
Z
= − cos(2t)e−st dt.
2 2 0
1 s2
Y (s) = − Y (s).
2 4
Solving for Y , we get that Y = 2/(s2 + 4). Using the L(y) notation,
we have that
2
L(sin 2t) = 2 .
s +4
Example 2: If y(t) is given in (6.6), find Y (s).
Answer: Using the additive property of integrals,
Z ∞
Y (s) = y(t)e−st dt
0
Z 2 Z ∞
−st
= y(t)e dt + y(t)e−st dt.
0 2
3 2 1 ∞
= − e−st + e−st
s t=0 s t=2
4 3
= − e−2s + .
s s
To guarantee that 1s e−st has a defined limit as t → ∞, it has been
assumed that Re(s) > 0.
Linear Operator
It states in (6.7) that L(3t − sin 2t) = 3L(t) − L(sin 2t). This is an
illustration of an essential property of the Laplace transform. Namely, if
c1 and c2 are constants, then
L c1 y1 + c2 y2 = c1 L(y1 ) + c2 L(y2 ). (6.8)
Another way to write this is to let y(t) = c1 y1 (t) + c2 y2 (t), in which case
Y (s) = c1 Y1 (s) + c2 Y2 (s), (6.9)
where Y1 and Y2 are the Laplace transforms for y1 and y2 , respectively.
Because the Laplace transform has this property, it is said to be a linear
operator. The usefulness of the linearity of the Laplace transform is why
it is listed first in Table 6.1.
It is worth pointing out that you know several other linear operators.
One is a matrix, because it satisfies A(c1y1 + c2 y2 ) = c1 Ay 1 + c2dAy2 . A
d
second is differentiation, as it satisfies dt c1 y1 (t) + c2 y2 (t) = c1 dt y1 (t) +
d
c2 dt y2 (t). You are probably wondering if it’s possible for mathematicians,
or even engineers, to write entire textbooks on linear operators. Well, a
Google book search will answer this question.
Using a Table
Table 6.1, and specifically Properties 7-10 in the table, make it easier to
find some of the more common Laplace transforms. You simply find the
function y(t) in the third column, and then its Laplace transform Y (s) is
in the second column. The next example illustrates how this is done.
Example 3 (version 2): If y(t) = 3t − sin 2t, find Y (s).
Answer: Using linearity, as expressed in (6.8),
Y (s) = L 3t − sin 2t = 3L(t) − L(sin 2t).
From Table 6.1, using Property 8 (with n = 1 and a = 0), and
Property 9 (with a = 0 and ω = 2),
3 2
Y (s) = 2
− 2 .
s s +4
It is possible to extend the usefulness of Table 6.1 by using some of the
properties of a Laplace transform. Exercise 4 considers one of particular
note.
Exercises
1. Find the Laplace transform of the following functions.
150 Chapter 6. Laplace Transform
3. One way to avoid using integration by parts is to use the formulas
cos x = 21 eix + e−ix and sin x = 2i1
eix − e−ix (see Section 3.4.1).
Use this, and Table 6.1, to find the Laplace transform of the following
functions:
6. Y (s + a) e−at y(t)
1 −as
7. se H(t − a) for a > 0
n!
8. (s+a)n+1 tn e−at for n = 0, 1, 2, 3, . . .
ω
9. (s+a)2 +ω 2 e−at sin(ωt)
s+a
10. (s+a)2 +ω 2 e−at cos(ωt)
h i
c + (ac + d)t eat if s1 = s2 = a
cs+d 1
h
at − (bc + d)ebt
i s1 = a
11. (s−s )(s−s ) a−b (ac + d)e if s2 = b
1 2
s = a + ib
h i
eat c cos(bt) + ac+db sin(bt) if s1 = a − ib
2
p(s)
12. (s−s )(s−s )···(s−s ) see (6.16)
1 2 n
14.
1 −as 1 n−1 e−b(t−a) H(t−a) for a > 0,
(s+b)n e (n−1)! (t−a) n = 1, 2, 3, . . .
Z t−a
1 −as
15. se Y (s) H(t − a) y(r)dr for a > 0
0
Table 6.1. Laplace and inverse Laplace transforms. The function H(x) is
defined in (6.13), and δ(t) is defined in Section 6.5.1. Also, recall that 0! = 1, and if
t > 0, then t0 = 1.
152 Chapter 6. Laplace Transform
The caveat here is that if Y = L(y), it is not always true that y = L−1 (Y ).
It is true for the above example, and this is because the original function
y(t) is continuous. What happens when y(t) has a jump discontinuity will
be discussed later.
In Section 6.4, the first differential equation we will solve using a
1 1
Laplace transform is y ′ + 3y = e2t . We will find that Y = s+3 (2 − 2−s ),
−1
and this will mean that to find y we will need to determine L (Y ). There
is a general formula for the inverse Laplace transform, which involves
a line integral in the complex plane. Although this can provide some
entertaining mathematical challenges, most find the inverse transform by
using tables. Table 6.1 is an example, and it is the one used in this text.
Note that the first six entries are general properties for the transform.
The first one listed is the linearity property, as given in (6.8). Writing it
in terms of the inverse transform, we have that
3 7s
Example: If Y (s) = s2 − s2 +25 , find y(t).
Answer: Using the linearity property,
3 7s
L−1 (Y ) = L−1 −
s2 s2 + 25
1 s
= 3L−1 2 − 7L−1 2 .
s s + 25
Therefore,
y(t) = 3t − 7 cos(5t).
6.2. Inverse Laplace Transform 153
then
4 3
Y (s) = − e−2s + .
s s
Except
for t = 2, L−1 (Y ) = y. At t = 2, the average in the jump in y(t)
is 2 y(2 ) + y(2− ) = 21 (3 − 1) = 1. Therefore, the inverse transform is
1 +
3 if 0 ≤ t < 2,
−1
L (Y ) = 1 if t = 2, (6.12)
−1 if 2 < t.
and this is shown in Figure 6.1. Note that this has built into its definition
the value at a jump that is needed for the inverse Laplace transform.
To rewrite (6.12) using H(x), since L−1 (Y ) involves a jump of −4 at
t = 2, then L−1 (Y ) = 3 − 4H(t − 2).
1
H(x)
0.5
0
-1 0 1
2 5 6
Example: If Y = s + s e−3s − s e−4s , find and then sketch y.
Answer: Using the linearity property,
2 5 6
L−1 (Y ) = L−1 + e−3s − e−4s
s s s
1 1 1
= 2L−1 + 5L−1 e−3s − 6L−1 e−4s .
s s s
From Property 8 in Table 6.1, with n = 0 and a = 0,
1
L−1 = 1.
s
From Property 7, with a = 3 and a = 4,
1 1
L−1 e−3s = H(t − 3), and L−1 e−4s = H(t − 4).
s s
Therefore,
y(t) = 2 + 5H(t − 3) − 6H(t − 4). (6.14)
For those who are picky about doing things correctly, there is a mild
case of notation abuse in the last example. Because the function y(t)
has a jump, and we only know its Laplace transform, it is not possible
to determine the value of y(t) at the jump. The function given in (6.14)
is the answer that is consistent with the formula determined using the
inverse Laplace transform. This situation will arise in this chapter any
time the function y(t) has a jump discontinuity.
6
y(t)
0
0 1 2 3 4 5 6
t
Exercises
1. Sketch the function for 0 ≤ t ≤ T , and then find its Laplace transform.
156 Chapter 6. Laplace Transform
3. Suppose that y(t) is periodic with period T > 0. So, y(t + T ) = y(t)
for all t ≥ 0.
a) Show that Z ∞
y(t)e−st dt = e−sT Y (s).
T
R∞ RT R∞
b) Writing 0 y(t)e−st dt = 0 y(t)e−st dt + T y(t)e−st dt, use the re-
sult from part (a) to show that
T
1
Z
L(y) = y(t)e−st dt.
1 − e−sT 0
4. The following functions are periodic with period T . Sketch the function
for 0 ≤ t ≤ 3T , and then use the result of Exercise 3(b) to find the
Laplace transform. Also, provide an explanation for where the name
of the wave comes from.
a) Square wave: T = 2, and y(t) = H(t) − H(t − 1), for 0 ≤ t < 2.
b) Sawtooth wave: T = 1, and y(t) = t, for 0 ≤ t < 1.
c) Triangle wave: T = 2, and y(t) = tH(t) − 2(t − 1)H(t − 1), for
0 ≤ t < 2.
d) Bang-bang wave: T = 2, and y(t) = H(t) − 2H(t − 1), for 0 ≤ t < 2.
6.3. Properties of the Laplace Transform 157
5. The floor function y(t) = ⌊t⌋ is the greatest integer less than or equal
to t. So, ⌊5.3⌋ = 5 and ⌊7.0⌋ = 7.
a) Writing ⌊t⌋ = t − g(t), what are: g(0), g(0.1), g(0.8), and g(1)?
b) Sketch g(t) for 0 ≤ t < 5. Use this to explain why g(t) is periodic.
c) Use the result from Exercise 3(b) to find L(⌊t⌋).
6. It is sometimes necessary to use a power series to determine an inverse
transform. For example, to use the geometric series (1 − z)−1 = 1 −
z + z 2 + · · · to write (1 − e−s )−1 = 1 − e−s + e−2s + · · · . In this problem
you are to use a Maclaurin series to find the inverse Laplace transform.
1
a) Y = s(1−e−s ) , use the geometric series for (1 − z)−1
1
b) Y = √ , use the series for (1 + z)−1/2
s 1+e−s
1p √
c) Y = s 1 + (1/s), use the series for 1 + z
In the last line above, interchanging the order of integration used the fact
that 0 < t < ∞, 0 < τ < t is equivalent to 0 < τ < ∞, τ < t < ∞. Now,
making the change of variables t = r + τ in the inner integral, we get that
Z ∞Z ∞
L(y) = g(r)v(τ )e−s(r+t) drdτ
Z0 ∞ 0 Z ∞
−sτ −sr
= v(τ )e g(r)e dr dτ = V (s)G(s).
0 0
Exercises
1. Find the Laplace transform in terms of Y (s).
a) L y ′ − 4y , where y(0) = 1
b) L 2y ′ + 7y , where y(0) = −2
s 1 s
b) (s2 +1)2 , taking V (s) = s2 +1 and G(s) = s2 +1
5
c) (s+1)(s2 +4)
1
d) (s2 +1)(s2 −1)
1
e) s3 (s2 +1)
3. Explain why Property 4 in Table 6.1 is a special case of the convolution
theorem.
Also,
∞ ∞
1
Z Z
2t 2t −st
L(e ) = e e dt = e(2−s)t dt = .
0 0 s−2
Since, using Property 9 in Table 6.1, L(sin t) = 1/(s2 + 1), then the
transformed problem is
1
(s2 + s − 2)Y − s − 2 = − .
s2 + 1
Solving for Y gives us
1 1
Y = − 2 . (6.24)
s − 1 (s + 1)(s2 + s − 2)
1
−1 As + B
Cs + D
L−1 = L + L −1
(s2 + 1)(s2 + s − 2) s2 + 1 s2 + s − 2
1 h i
= A cos t + B sin t + (C + D)et + (2C − D)e−2t .
2
(6.25)
Therefore, the solution is
1 3 1 5
y= cos t + sin t + e−2t + et .
10 10 15 6
Once you know the transfer function, then a particular solution of the
differential equation can be written down using the convolution theorem.
Namely, using (6.21), a particular solution is
Z t
yp (t) = h(t − τ )f (τ )dτ , (6.26)
0
1 t h 3(t−τ ) i√
Z
−t
y(t) = e + e − e−t+τ τ dτ.
4 0
Exercises
1. Use the Laplace transform to find the solution of the IVP.
a) 2y ′ + y = 1, y(0) = 2
b) 3y ′ = −y + e−t , y(0) = 21
c) y ′′ + y ′ − 2y = 0, y(0) = 0, y ′ (0) = −1
d) y ′′ − 6y ′ + 9y = 0, y(0) = 0, y ′ (0) = 2
e) 5y ′′ − y ′ = 0, y(0) = −1, y ′ (0) = −1
f) 4y ′′ + y = 0, y(0) = −1, y ′ (0) = −1
g) y ′′ − 2y ′ + 2y = 0, y(0) = −1, y ′ (0) = −1
h) y ′′ + 2y ′ + 5y = 0, y(0) = 0, y ′ (0) = −6
2. Use the Laplace transform to find the solution of the IVP.
a) y ′′ + y ′ − 2y = 12t, y(0) = 0, y ′ (0) = 0
b) y ′′ + 4y = 8t2 , y(0) = 0, y ′ (0) = 0
c) y ′′ − y ′ = 2 sin t, y(0) = 0, y ′ (0) = 0
d) y ′′ + 3y ′ = 3t + 1, y(0) = 0, y ′ (0) = 0
e) y ′′ − 2y ′ + 5y = 5 − 4e−t , y(0) = 0, y ′ (0) = 0
3. For the following, find the transfer function H(s) and then write down
the resulting particular solution. You do not need to evaluate the
integral.
a) y ′ + 3y = ln(1 + 3t)
√
b) y ′′ + 9y = 1 + t
c) 2y ′′ + 3y ′ − 2y = 1/(1 + t)
d) y ′′ + 2y ′ + 5y = sin(1 + t2 )
4. Proceeding as in Example 3, find the solution of the following IVPs.
a) y ′ + 3y = ln(1 + 3t), where y(0) = 1
√
b) y ′′ + 9y = 1 + t, where y(0) = 1 and y ′ (0) = 0
c) 2y ′′ + 3y ′ − 2y = 1/(1 + t), where y(0) = 2 and y ′ (0) = −3
d) y ′′ + 2y ′ + 5y = sin(1 + t2 ), where y(0) = 0 and y ′ (0) = 2
164 Chapter 6. Laplace Transform
= (s2 + 3s + 2)Y − s − 2.
R3
Also, L(f ) = 0 2e−st dt = 2 1−e−3s /s. Consequently, from (6.27),
we have that
2
(s + 1)(s + 2)Y = s + 2 + 1 − e−3s ,
s
which means that
1 2
−3s
Y = + 1−e . (6.28)
s + 1 s(s + 2)(s + 1)
To determine the inverse transform, using Property 8 from Table
6.1, L−1 (1/(s + 1)) = e−t . Also, from Property 11,
1
L−1 = e−t − e−2t .
(s + 2)(s + 1)
Consequently, using Properties 5 and 15 (respectively),
2
L−1 1 − e−3s
s(s + 2)(s + 1)
−1
2
−1
2 −3s
=L −L e
s(s + 2)(s + 1) s(s + 2)(s + 1)
Z t Z t−3
= 2 (e−r − e−2r )dr + 2H(t − 3) (e−r − e−2r )dr
0 0
= −2e−t + e−2t + 1 + H(t − 3)(1 − 2e3−t + e−2t+6 ).
The resulting solution is therefore
y = 1 + e−2t − e−t − 1 + e−2(t−3) − 2e−(t−3) H(t − 3).
6.5. Solving Equations with Non-Smooth Forcing 165
where Z t0 +ε
D= d(r)dr.
t0 −ε
We are assuming that the forcing interval is very short, but D is large
enough to be meaningful. To put this in physical terms, it is as if we are
hitting the system with a hammer.
There is a mathematical idealization for a concentrated force that
makes solving the problem easier than trying to use a formulation as in
(6.29). This is done by introducing what is known as the delta function.
Delta Function. The delta function δ(t) is defined to have the following
properties:
1. Given any t0 ,
δ(t − t0 ) = 0, when t 6= t0 . (6.31)
166 Chapter 6. Laplace Transform
and
t0 b
1
Z Z
δ(t − t0 )g(t)dt = δ(t − t0 )g(t)dt = g(t0 ) . (6.33)
a t0 2
y ′ = Dδ(t − t0 ),
To evaluate this, first note that if 0 ≤ t < t0 , then from (6.31), y(t) = 0.
If t = t0 , then from (6.33), y(t0 ) = D/2. Lastly, when t0 < t, then from
(6.32), y(t) = D. Consequently, the solution is
0
if 0 ≤ t < t0 ,
1
y=
2D if t = t0 , (6.34)
D if t0 < t.
Except for the very small time interval t0 − ε < t < t0 + ε, this solution
is the same as the one in (6.30). Moreover, the above solution is consis-
tent with what is obtained using the line integral formula for the inverse
Laplace transform.
The rationale for the stated properties of the delta function can be
explained by considering the case of when d is constant. The assumption is
that the total force D, what is known as the impulse, remains fixed as the
time interval decreases (see Figure 6.3). This requires that d = D/(2ε).
In other words, the magnitude of the force increases as the time interval
decreases. Consequently, in the limit, the forcing is zero if t 6= t0 and it is
infinite at t = t0 . This explains (6.31), and it also explains why you will
see the statement that δ(0) = ∞. This limit can also be used to explain
(6.32). Finally, it is being assumed that the impulse forcing is symmetric
about t0 , as it is in the case of when d is constant, and this gives us (6.33).
6.5. Solving Equations with Non-Smooth Forcing 167
Figure 6.3. A fixed impulse, applied over the time interval t0 − ε < t < t0 + ε,
used to explain the stated properties of the delta function.
= e−as .
Rt
Example 4: Evaluate −∞ δ(r)dr, for −∞ < t < ∞.
Answer: This can be answered by considering three cases. First, if
t < 0, then δ(r) = 0 for −∞ < r ≤ t, and we conclude the integral
is zero. If t > 0, then from (6.32) with g(t) = 1, the integral is equal
to one. Finally, when t = 0, from (6.33) we get the value of 1/2.
Therefore, we have that
Z t
H(t) = δ(r)dr.
−∞
H ′ (t) = δ(t).
Mathematical Tidbits
As you likely noticed, δ(t) is not actually a function. The more accurate
statement is that it is a distribution, or a generalized function. There are
various ways to obtain a mathematically rigorous definition of δ(t), using
limits or test functions. How limits are used was explained very briefly
earlier. This will not be pursued any further, but the question does arise
as to what is permitted when using the delta function. As demonstrated
in Example 2, linear combinations of delta functions are allowed. It is
also possible to both differentiate and integrate a delta function. What
should be avoided is using a discontinuous g(t) in (6.32), and Exercise 4 is
an example why. Also, what is not allowed, generally, involves nonlinear
operations. So, expressions such as δ(t − 1)δ(t − 2), δ(t − 1)/δ(t − 2), and
sin(δ(t−1)) are not allowed. If you are interested in the various properties
of the delta function, you might look at its Wikipedia page.
The nonstandard nature of the delta function amplifies a complication
with the Laplace transform at t = 0 that needs to be mentioned. It is
not uncommon in certain applications to use the forcing function f (t) =
δ(t), which means that it is located at t = 0. This puts it at the lower
point of integration for the Laplace transform.
The resulting integral can
be evaluated using (6.33), giving L δ(t) = 1/2. However, it is almost
universally stated that L δ(t) = 1. One way to explain this involves
continuity, in the sense that this is what you obtain from Property 13, in
+
Table 6.1, when letting a → 0 . On the other hand, one gets L δ(t) = 0
when letting a → 0− . This has lead those determined to obtain L δ(t) =
1 to find some rather creative ways to redefine the Laplace transform.
What this involves is not considered here, but if you are interested in
learning more about this issue, you should consult Hoskins [2009].
Exercises 169
Exercises
1. Use the Laplace transform to find the solution of the IVP.
a) y ′ + 4y = 3H(t − 1), y(0) = 1
b) 2y ′ − y = 1 − H(t − 4), y(0) = −1
c) y ′ + y = 2δ(t − 3), y(0) = −1
d) y ′ − 4y = 2H(t − 2) − δ(t − 1), y(0) = 0
e) y ′′ − y ′ − 6y = 3H(t − 5), y(0) = 0, y ′ (0) = 0
f) y ′′ + 4y = 3H(t − 4) − 3H(t − 2), y(0) = 0, y ′ (0) = 0
g) y ′′ − 4y ′ = 3δ(t − 1), y(0) = 0, y ′ (0) = 0
h) y ′′ + y = δ(t − 3) − 2δ(t − 2), y(0) = 0, y ′ (0) = 0
2. There are usually multiple ways to find an inverse transform, and this
exercise illustrates this by reconsidering (6.28).
a) Using partial fractions, the assumption is
2 A B C
= + + .
s(s + 2)(s + 1) s s+2 s+1
Find A, B, and C, and then determine the inverse transform.
b) Find the inverse transform of the function in part (a) but use Prop-
erty 12, with n = 3.
3. Show that the following identities hold for the delta function. Do
this by showing that when the left and right sides of the equation are
inserted into (6.31)-(6.33), that they produce the same result.
a) δ(a(t − t0 )) = a1 δ(t − t0 ), for a > 0
b) δ(t0 − t) = δ(t − t0 )
c) If g(t) is continuous, then g(t)δ(t − t0 ) = g(t0 )δ(t − t0 ) .
4. In quantum physics there are occasions when the coefficients of the
differential equation contain delta functions. The point of this exercise
is to demonstrate that care is needed in such situations.
a) Consider the problem of solving
y ′ (t) = δ(t − t0 )y(t), for t > 0,
where t0 > 0 and y(0) = 1. Using separation of variables, and
Example 4, find the solution. Make sure to determine its value for
0 ≤ t < t0 , for t = t0 , and for t0 < t. For the record, this is the
correct solution of this problem.
b) By simply integrating the differential equation in part (a), and then
using the initial condition, one gets that
Z t
y(t) = 1 + y(r)δ(r − t0 )dr.
0
170 Chapter 6. Laplace Transform
Not thinking too hard about the situation, and using (6.31)-(6.33),
explain how you might conclude that
1
if 0 ≤ t < t0 ,
y= 2 if t = t0 ,
3 if t0 < t.
This differs from the solution for part (a). Where is the error made
in the derivation of the above solution?
x′ = ax + by + f (t), (6.35)
′
y = cx + dy + g(t), (6.36)
sX − x0 = aX + bY + F, (6.37)
sY − y0 = cX + dY + G, (6.38)
sX − x0 = AX + F, (6.39)
x′ = x − y,
y ′ = 4x − 2y,
sX − 1 = X − Y,
sY + 1 = 4X − 2Y.
y = x − x′
−t/2 11
=e − cos(ωt) + sin(ωt) .
2ω
x′ = 3x − 6y + f (t),
y ′ = x − 4y + g(t),
where x(0) = 0 and y(0) = 0. Also, f (t) and g(t) are continuous
functions.
Answer: From (6.37) and (6.38) the transformed equations are
sX = 3X − 6Y + F,
sY = X − 4Y + G.
and
s − 3 1 −3t
2t
L−1 = 6e − e .
s2 + s − 6 5
So, using the convolution theorem, which is Property 2 of Table 6.1,
1 Z t 1
−1 2(t−r) −3(t−r)
L F (s) = e − e f (r)dr,
s2 + s − 6 0 5
and
t
s−3 1 −3(t−r)
Z
L −1
G(s) = 6e − e2(t−r) g(r)dr.
s2 + s − 6 0 5
Therefore, the solution is
1 t 2(t−r)
Z
− e−3(t−r) f (r)dr
y(t) = e
5 0
1 t
Z
6e−3(t−r) − e2(t−r) g(r)dr.
+
5 0
To find x you can either find the inverse transform for X, or you
can use the second differential equation (similar to what was done
in the previous example).
This requires finding the inverse matrix and then trying to determine the
inverse transform of the resulting formula for X. There are ways this can
be done, such as using a geometric series expansion for (sI − A)−1 , but
how this is carried out is beyond the purview of this textbook. Those
interested might want to look at Friedland [2005] and Cohen [2007].
Exercises 173
Exercises
1. Use the Laplace transform to find the solution of the IVP, with
4
x(0) = .
−1
! ! !
′
−1 6 ′
3 1 ′
2 0
a) x = x c) x = x e) x = x
1 0 1 3 −1 2
! ! !
′
0 41 ′
2 1 ′
1 1
b) x = x d) x = x, f) x = x
1 0 6 3 −4 1
x′ = ax + by,
y ′ = cx + dy,
! !
−1 6 2 0
a) x′ = x c) x′ = x
1 0 −1 2
! !
1
0 4 1 −4
b) x′ = x d) x′ = x
1 0 1 1
Chapter 7
Partial Differential
Equations
Each of these PDEs is linear and homogeneous. Also, the advection equa-
tion is first order, while the other two are second order.
Subscripts are used in the above PDEs to indicate partial differentia-
tion. There are two other ways this can be done that are very common.
First, there is the form used in calculus, and examples are
175
176 Chapter 7. Partial Differential Equations
where
u(0) = 1, (7.2)
and
u(2) = −3. (7.3)
This is called a boundary value problem (BVP), and it consists of a
differential equation and two boundary conditions, one at each end
of the spatial interval. Because this involves a linear differential equation
with constant coefficients, the methods developed in Chapter 3 can be
used to solve it. So, assuming that u = erx , and then substituting this
into the differential equation (7.1), you obtain the characteristic equation
r2 = 4. The two solutions are r1 = −2 and r2 = 2, which means that the
general solution of (7.1) is
u = c1 e−2x + c2 e2x .
7.2. Boundary Value Problems 177
e4 + 3 e−4 + 3
c1 = and c2 = − .
e4 − e−4 e4 − e−4
2Ax − 3A + 2B = 4x.
u = 2x + 3 + c1 ex + c2 e2x .
11
x 2x
u = 2x + 3 + e − e .
e4 (e4 − 1)
Example 2: Show that u′′ + u = 0, where u(0) = 1 and u(π) = −3, has
no solution.
where
u(0) = 0, (7.5)
and
u(1) = 0. (7.6)
The function u = 0 is a solution, but what we want to know is whether
there are nonzero solutions. To be specific, is it possible to find values
of the constant λ so there are solutions that are not identically zero?
This is the same question asked when solving the eigenvalue problem
Aa = λa. In other words, finding u(x) and λ is an eigenvalue problem.
In this context, the u’s are called eigenfunctions, and the λ’s are the
eigenvalues. A distinctive difference from the matrix eigenvalue problem
is that there can be an infinite number of eigenvalues for an eigenvalue
BVP.
Finding λ and u is not hard. As usual, assuming that u = erx , then
the characteristic
√ equation coming from (7.4) is r2 − λ = 0. This means
that r = ± λ. Assuming λ is a real number then we have the following
three cases:
√ √
λ > 0 : In this case, the general solution is u = ae λx +be− λx
√ . To satisfy
√
u(0) = 0 we need a+b = 0, and for u(1) = 0√we need√ae λ +be− λ =
0.√ So, b = √
−a, and this means that a(e λ − e− λ ) = 0. Since
e λ 6= e− λ when λ > 0, the conclusion is that a = 0, and this
means we just get the zero solution.
Rayleigh Quotient
It is possible to show that the eigenvalues for the above BVP must be
negative, without having to first derive the formula for them. This can
be done using what is called the Rayleigh quotient, and this is explained
in Exercise 4. In fact, the steps in this exercise can be modified to also
prove that the eigenvalues must be real-valued, which is an assumption
we made in solving the eigenvalue problem.
The Rayleigh quotient is more than a theoretical tool as it plays an im-
portant role when studying mechanical vibrations as well as when finding
quantum energy levels. It is also used extensively in scientific computing
when solving eigenvalue problems.
Exercises
1. Solve the given BVP.
a) u′′ − 4u = 0, for 0 < x < 2; u(0) = 0 and u(2) = 1.
b) u′′ + u = 0, for 0 < x < 1; u(0) = 0 and u(1) = −1.
c) u′′ + u′ + u = 0, for 0 < x < 1; u(0) = 0 and u(1) = 1.
d) u′′ − u = 5, for 0 < x < 2; u(0) = 0 and u(2) = 0.
e) u′′ + u′ = x, for 0 < x < 1; u(0) = 0 and u(1) = 0.
2. Show that the given BVP has no solution.
180 Chapter 7. Partial Differential Equations
The Rayleigh quotient for this problem is obtained when you solve
the above equation for λ.
b) Use part (a) to explain why, given an eigenfunction u(x), that the
associated eigenvalue must be negative.
c) The fundamental eigenfunction corresponds to the case of n = 1 in
(7.7). Taking b1 = 1, sketch u1 (x) for 0 ≤ x ≤ 1. On the same axes,
also sketch w(x) = 4x(1 − x).
d) Part (c) shows that w(x) can be used as an approximation of u1 (x).
Use w in the Rayleigh quotient to obtain an approximation for λ1 .
u(0, t) = 0, (7.10)
and
u(L, t) = 0. (7.11)
For the initial condition, it is assumed that
Substituting this into the PDE (7.9) gives DF ′′ (x)G(t) = F (x)G ′ (t).
Separating variables yields
F ′′ (x) G ′ (t)
D = . (7.14)
F (x) G(t)
Now comes the key observation. The only way a function of x can equal
a function of t, since x and t are independent, is that the function of x is
a constant, the function of t is a constant, and the constants are equal.
In other words, there is a constant λ so that
F ′′ (x)
D = λ,
F (x)
and
G ′ (t)
= λ.
G(t)
182 Chapter 7. Partial Differential Equations
and
G ′ (t) = λG(t) . (7.16)
The λ appearing here is called, not surprisingly, the separation con-
stant.
or equivalently
√ √ √
π D 3π D 3π D
k= , , ,.... (7.19)
L L L
7.3. Separation of Variables 183
The conclusion is that the not identically zero solutions of (7.17) and
(7.18) are nπx
Fn (x) = bn sin , (7.20)
L
and nπ 2
λn = −D , (7.21)
L
for n = 1, 2, 3, . . .. Also, bn is an arbitrary constant.
This is the equation that is used to determine the bn ’s. However, the
left-hand-side is an example of what is known as a Fourier series. More
184 Chapter 7. Partial Differential Equations
2 L nπx
Z
bn = g(x) sin dx. (7.25)
L 0 L
7.3.6 Examples
Example 1: Suppose that D = 1, L = 2, and g(x) = 3 sin(πx). In this
case, from (7.21), λn = −(nπ/2)2 , and the resulting general solution
(7.23) is
∞ nπx
2 2
X
u(x, t) = bn e−n π t/4 sin .
2
n=1
3
t=0
t = 0.05
Solution
t = 0.15
0
-3
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
x-axis
or equivalently
πx 3πx
b1 sin +b2 sin(πx)+b3 sin +b4 sin(2πx)+· · · = 3 sin(πx).
2 2
To satisfy this equation, take b2 = 3 and set all the other bn ’s to
zero. Therefore, the solution is
2
u(x, t) = 3e−π t sin(πx). (7.26)
This solution is shown in Figure 7.1, both as time slices and as the
solution surface for 0 ≤ t ≤ 0.24.
1 t=0
t = 0.002
Solution
t = 0.03
0.5 t = 0.2
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
x-axis
Exercises
1. You are to find the solution of the diffusion problem for the following
initial conditions. Assume that L = 1 and D = 3. Note that you
should be able to answer this question without using integration.
a) g(x) = −4 sin(5πx).
b) g(x) = 6 sin(11πx).
c) g(x) = sin(πx) + 8 sin(4πx) − 10 sin(7πx).
d) g(x) = − sin(3πx) + 7 sin(8πx) + 2 sin(15πx).
e) g(x) = 4 sin(2πx) cos(πx).
Exercises 187
2. You are to find the solution of the diffusion problem for the following
initial conditions. Assume that L = 2 and D = 4.
a) g(x) = 1 −1 if 0 ≤ x ≤ 1
d) g(x) =
0 otherwise
b) g(x) = 2 + x
1 if 0 ≤ x ≤ 13
c) g(x) = cos(πx) e) g(x) =
2 otherwise
if 12 ≤ x ≤ 3
1 2
where u(0, t) = 0, u(2, t) = 0, and u(x, 0) =
0 otherwise.
5. Find the general solution of the following.
a) uxx = ut , for 0 < x < 1, with the boundary conditions u(0, t) = 0
and ux (1, t) = 0.
b) 4uxx = ut , for 0 < x < 1, with the boundary conditions ux (0, t) = 0
and u(1, t) = 0.
c) (1 + t)∂x2 u = ∂t u, for 0 < x < 1, with the boundary conditions
u(0, t) = 0 and u(1, t) = 0.
d) uxx = ut + e−t u, for 0 < x < 1, with the boundary conditions
u(0, t) = 0 and u(1, t) = 0.
6. Find the solution of the problem for the given initial condition.
a) Exercise 5(a), with u(x, 0) = 3 sin πx 9πx
2 −7 sin 2 .
b) Exercise 5(b), with u(x, 0) = −5 cos 3πx 11πx
2 − 2 cos 2 .
c) Exercise 5(c), with u(x, 0) = 14 sin(10πx) + 30 sin(18πx).
d) Exercise 5(d), with u(x, 0) = −24 sin(3πx) − 12 sin(15πx).
7. Find the resulting ODEs obtained using separation of variables on the
given PDE.
a) (1 + x)uxx + tu = 7ut , assuming u = F (x)G(t)
b) r2 urr + rur + uθθ = 0, assuming u = R(r)Θ(θ)
c) ∂x (ex ∂x u) = (1 + x2 )∂t u, assuming u = F (x)G(t)
d) uzz + 3zuz = uyy + 9u, assuming u = Z(z)Y (y)
e) u2x + u2t = e−t u2 , assuming u = F (x)G(t)
188 Chapter 7. Partial Differential Equations
To illustrate how (7.30) is used, suppose we want to find the value for,
say, b7 . Multiplying (7.29) by sin(7πx/L), and then integrating yields
Z L 7πx ∞ Z L
X 7πx nπx
g(x) sin dx = bn sin sin dx.
0 L 0 L L
n=1
According to (7.30), all of the integrals on the right are zero except when
n = 7. Consequently,
Z L 7πx L
g(x) sin dx = b7 ,
0 L 2
7.4. Sine and Cosine Series 189
or equivalently
L
2 7πx
Z
b7 = g(x) sin dx.
L 0 L
A similar result is obtained for the other bn ’s, and the resulting formula
is
2 L nπx
Z
bn = g(x) sin dx. (7.31)
L 0 L
Sine Series Convergence Theorem. Assume that g(x) and g ′ (x) are
piecewise continuous for 0 ≤ x ≤ L, and the bn ’s are given in (7.31).
For 0 < x < L: If g(x) is continuous at x, then
∞
X nπx
g(x) = bn sin , (7.32)
L
n=1
In words, the theorem states that the sine series equals the function g(x)
at points in the interval where g(x) is continuous, and it equals the average
in the jump of g(x) at a jump discontinuity. At the endpoints, no matter
what the value of g(0) or g(L), the series sums to zero.
7.4.3 Examples
Finding a sine series is rather uneventful as it is simply a matter of eval-
uating the given formulas. The only concern is how hard it is to evaluate
the integrals to find the coefficients. So, in the examples below, a more
practical question is also considered. Namely, how many terms of the
series do you have to add together to obtain an accurate approximation
of the function g(x)? As will be seen, the answer depends on whether
the function is continuous, and whether it has the right values at the
endpoints.
190 Chapter 7. Partial Differential Equations
g(x)
1
0
0 1 2 3
x-axis
1
g (x)
0.5
0
0 1 2 3
x-axis
2
Sine Series
0
0 1 2 3
x-axis
Figure 7.3. The functions g(x), g ′ (x), and the function that the sine series
sums to for Example 1.
Sketch g(x) and g ′ (x) for 0 ≤ x ≤ 3, and use this to explain why
they are piecewise continuous. Also, sketch the function that the
sine series for g(x) converges to for 0 ≤ x ≤ 3.
Answer: The functions g(x) and g ′ (x) are shown in Figure 7.3.
Although g(x) is not continuous at x = 2, and g ′ (x) is not defined
there, the left and right limits of both functions are defined and finite
at that point. Therefore, both functions are piecewise continuous
for 0 ≤ x ≤ 3. As for the sine series, for 0 < x < 3, it equals
g(x) except at the discontinuity, where it sums to the average in the
jump. So, at x = 2, it converges to 21 [g(2+ ) + g(2− )] = 21 (1 + 2) = 23 .
Finally, at x = 0, and at x = 3, the series sums to zero. The sketch
of the resulting function is given in Figure 7.3.
√
Example 2: Taking L = 5, suppose g(x) = x. Are g(x) and g ′ (x)
piecewise continuous for 0 ≤ x ≤ 5?
Answer: The function g(x) is continuous for 0 ≤ x ≤ 5 (and it is,
7.4. Sine and Cosine Series 191
√
therefore, piecewise continuous). Its derivative g ′ (x) = 1/(2 x) is
continuous for 0 < x ≤ 5, but g ′ (0+ ) = ∞. Because this limit is not
finite, g ′ (x) is not piecewise continuous for 0 ≤ x ≤ 5.
∞
X 9
g(x) = sin(nπ/3) sin(nπx), for 0 ≤ x ≤ 1. (7.34)
π 2 n2
n=1
1 1
g(x)
g(x)
0.5 0.5
N=1 N=3
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x-axis x-axis
1 1
g(x)
g(x)
0.5 0.5
N=9 N = 27
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x-axis x-axis
Figure 7.4. Comparison between the function g(x) in Example 3, shown with
the dashed blue curve, and the sine series approximation in (7.35), shown using a solid
red curve.
192 Chapter 7. Partial Differential Equations
1
Sine Series
0.5
0
0 0.25 0.5 0.75 1
x-axis
Figure 7.5. The function the sine series in Example 4 converges to for 0 ≤ x ≤ 1.
7.4. Sine and Cosine Series 193
1 N=4 1 N = 20
g(x)
g(x)
0.5 0.5
0 0
1 N = 40 1 N = 200
g(x)
g(x)
0.5 0.5
0 0
Figure 7.6. Comparison between the function g(x) for Example 4, shown
with the blue curve, and the sine series approximation in (7.37), shown using the red
curve.
The convergence theorem for this is very similar to the one for the sine
series. First, the needed integration formula is, if m and n are integers,
L if m = n = 0,
Z L nπx mπx
L
cos cos dx = if m = n 6= 0, (7.39)
0 L L
2
0 if m 6= n.
In words, the theorem states that the cosine series equals the function
g(x) at points in the interval where g(x) is continuous, and it equals the
average in the jump of g(x) at a jump discontinuity. At the endpoints, it
sums to the respective limit of g(x) at the endpoint.
7.4. Sine and Cosine Series 195
2
Cosine Series
1.5
1
0 0.5 1
x-axis
Figure 7.7. The function the cosine series in Example 5 converges to for 0 ≤ x ≤ 1.
196 Chapter 7. Partial Differential Equations
2 2
g(x)
g(x)
1.5 1.5
N=4 N = 20
1 1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x-axis x-axis
2 2
g(x)
g(x)
1.5 1.5
N = 40 N = 200
1 1
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
x-axis x-axis
Figure 7.8. Comparison between the function g(x) for Example 5, shown
with the blue curve, and the cosine series approximation in (7.44), shown using the red
curve.
7.4.5 Differentiability
In using a sine or cosine series when solving a PDE, it is implicitly assumed
you can differentiate the series term-by-term. What this means is that it
is assumed that
∞ ∞
d X X d
pn (x) = pn (x).
dx dx
n=1 n=1
With this in mind, in Example 4, if you try this with (7.36), you get
∞
X
g ′ (x) =
2 1 − cos(nπ/4) cos(nπx). (7.45)
n=1
P
As you should recall, if an infinite series an converges, then it must be
true that an → 0 as n → ∞. The above series for g ′ (x) does not satisfy
this condition, and therefore it does not converge. In other words, you
can not differentiate (7.36) term-by-term. In contrast, for Example 3 you
can differentiate the series term-by-term. The theorem that explains this
states that if g(x) is continuous, and g ′ (x) is piecewise continuous, for
0 ≤ x ≤ L, then you can differentiate the cosine series term-by-term, but
to do this for a sine series you need an additional assumption [Tolstov and
Silverman, 1976]. An easy to use version of the needed assumption is that
g(0) = g(L) = 0. This holds for Example 3, and that is why term-by-term
7.4. Sine and Cosine Series 197
differentiation can be done with that sine series. For both the cosine and
sine series, if g(x) is not continuous, then term-by-term differentiation is
not possible without additional assumptions. Those interested in pursuing
this issue a bit further should look at Exercise 11.
The situation for term-by-term integration is better. Specifically, if
g(x) satisfies the requirements of the convergence theorem, its sine, and
cosine, series can be integrated term-by-term.
The next question is whether the potential non-differentiability of a
sine series means that we can not use it to solve the diffusion equation.
To explain why this is not a problem, consider the solution (7.43), which
is
∞
2 π2 t
X
u(x, t) = bn e−n sin(nπx).
n=1
Exercises
1. Sketch the graph of f (x) for 0 ≤ x ≤ 1. Also, determine whether f (x)
is continuous, piecewise continuous, or neither for 0 ≤ x ≤ 1.
1
1 if 0 ≤ x ≤ 2
0 if x = 0
1 3
a) f (x) = 2x if 2 < x ≤ 4 c) f (x) = ln x if 0 < x < 1
3
3
if 4 < x ≤ 1
1 if x = 1
2
( (
1 if x = 14 , 21 , 43 , 1 0 if 0 ≤ x ≤ 12
b) f (x) = d) f (x) = 1 1
2 otherwise 2x−1 if 2 < x ≤ 1
2. Assuming that L = 2, explain why g(x) does not satisfy the conditions
stated in the Sine Series Convergence Theorem.
1
a) g(x) = x1/3 b) g(x) = tan x c) g(x) = x2 +4x−1
u(0, t) = 0, (7.47)
and
u(L, t) = 0. (7.48)
For the initial conditions, it is assumed that
and
ut (x, 0) = h(x), for 0 < x < L, (7.50)
where g(x) and h(x) are given functions. To avoid the complication with
differentiability, as described in Section 7.4.5, it is assumed that g(x)
and h(x) are smooth functions that satisfy the boundary conditions, and
g ′′ (0) = g ′′ (L) = 0.
As with the diffusion problem, separation of variables will be used to
find the general solution of the PDE and boundary conditions. After that,
the initial conditions will be satisfied. Also, you should notice, as with the
diffusion problem, the PDE and boundary conditions are homogeneous.
This is required for separation of variables to work.
Assuming
u(x, t) = F (x)G(t), (7.51)
and then substituting this into the PDE gives us
F ′′ (x) G ′′ (t)
c2 = . (7.52)
F (x) G(t)
and
G ′′ (t) = λG(t) . (7.54)
and nπ 2
λn = −c2 , (7.58)
L
for n = 1, 2, 3, . . .. Also, b̄n is an arbitrary constant.
Finding G(t)
We have shown that for any given n, the function un (x, t) = Fn (x)Gn (t)
is a solution of the PDE that satisfies the boundary conditions. The re-
sulting general solution, that satisfies the PDE and boundary conditions,
is, therefore,
X∞
u(x, t) = un (x, t),
n=1
or equivalently
∞
X nπx
u(x, t) = an cos(ωn t) + bn sin(ωn t) sin , (7.61)
L
n=1
This is the same problem we had in Section 7.3.5, except that the
coefficient is being denoted as Bn instead of bn . So, from (7.25),
L
2 nπx
Z
Bn = h(x) sin dx.
L 0 L
7.5.1 Examples
Example 1: Suppose that c = 1, L = 2, g(x) = 3 sin(πx), and h(x) = 0.
In this case, from (7.60), ωn = nπ/2. The resulting general solution
(7.61) is
∞ h
X nπ nπ i nπx
u(x, t) = an cos t + bn sin t sin .
2 2 2
n=1
3
Solution
-3
0 0.5 1 1.5 2
x-axis
Figure 7.9. Solution of the wave equation in Example 1. Shown is the solution
surface as well as the solution profiles at specific time values.
where
cnπ
ωn = . (7.69)
L
The expression in the square brackets is a periodic function of t, with
period 2π/ωn . In this context, sin(nπx/L) is called a natural mode for
the problem, having natural frequency ωn . The resulting solution in (7.68)
corresponds to what is called a standing wave. So, the curves shown in
the lower plot in Figure 7.9 are plots of a standing wave in the case of
when n = 2. It is also possible to have traveling wave solutions, similar
to waves on a lake or ocean. If you want to learn about traveling waves,
you might look at Strauss [2007] or Holmes [2019].
Exercises
1. You are to find the solution of the wave equation problem for the
following initial conditions. Assume that L = 1 and c = 4. Note that
you should be able to answer this question without using integration.
a) g(x) = sin(3πx), and h(x) = 0
b) g(x) = 0, and h(x) = −2 sin(8πx)
c) g(x) = − sin(πx) + 4 sin(3πx), and h(x) = −3 sin(5πx)
d) g(x) = 5 sin(7πx), and h(x) = 2 sin(8πx) + 3 sin(12πx)
e) g(x) = 2 sin(2πx) cos(πx), and h(x) = −2 sin(8πx)
f) g(x) = 3 cos(2πx − π2 ), and h(x) = −3 cos(7πx) sin(2πx)
2. Find the general solution of the following.
a) uxx = utt , for 0 < x < 1, with the boundary conditions u(0, t) = 0
and ux (1, t) = 0.
b) 4uxx = utt , for 0 < x < 1, with the boundary conditions ux (0, t) = 0
and u(1, t) = 0.
c) uxx = 4utt , for 0 < x < 1, with the boundary conditions u(0, t) =
u(1, t) and ux (0, t) = ux (1, t).
d) uxx = utt +ut , for 0 < x < 1, with the boundary conditions u(0, t) =
0 and u(1, t) = 0. This is an example of what is called a damped
wave equation.
3. Solve
∂2u ∂2u
0 < x < 1,
4 2 = 2 , for
∂x ∂t 0 < t,
where u(0, t) = 0, u(1, t) = 0, u(x, 0) = 0, and ut (x, 0) = x(1 − x).
how to find the solution when the boundary conditions are inhomoge-
neous, and have the form
u(0, t) = α, (7.70)
and
u(L, t) = β, (7.71)
where α and β are constants. The method used to find the solution is to
write it as
u(x, t) = w(x) + v(x, t),
where we pick w(x) so it satisfies the given boundary conditions. In other
words, so that w(0) = α and w(L) = β. Pretty much any smooth function
can be used, but it makes things easier if w comes from the steady state
equation. What this entails is explained below.
d2 w
= 0, for 0 < x < L,
dx2
where, from (7.70) and (7.71), w(0) = α and w(L) = β. The resulting
solution is
β−α
w(x) = α + x.
L
v(0, t) = 0, (7.74)
and
v(L, t) = 0. (7.75)
206 Chapter 7. Partial Differential Equations
Finally, if the initial condition is u(x, 0) = g(x), then the resulting initial
condition for v is
β−α
v(x, 0) = g(x) − α − x, for 0 < x < L. (7.76)
L
The above problem for v(x, t) has the same form as the one for u(x, t),
as given in (7.9)-(7.12), except for a slightly different looking initial condi-
tion. Consequently, we can use the solution as given in (7.23) and (7.25)
if we make the appropriate adjustments. In particular,
X∞ nπx
v(x, t) = bn eλn t sin ,
L
n=1
where
L
2 β − α nπx
Z
bn = g(x) − α − x sin dx, (7.77)
L 0 L L
and nπ 2
λn = −D .
L
7.6.3 Summary
We have shown that the solution of
∂2u
∂u 0 < x < L,
D 2 = , for
∂x ∂t 0 < t,
where u(0, t) = α, u(L, t) = β, and u(x, 0) = g(x), is
∞
β−α X nπx
u(x, t) = α + x+ bn eλn t sin , (7.78)
L L
n=1
Exercises
1. You are to find the solution of the diffusion equation 4uxx = ut for the
given boundary and initial conditions. Assume that L = 1.
a) u(0, t) = 1, u(1, t) = −1, and u(x, 0) = 0.
b) u(0, t) = 2, u(1, t) = −5, and u(x, 0) = 2.
c) u(0, t) = −4, u(1, t) = 1, and u(x, 0) = x.
2. Find the steady state solution of the following problems.
a) uxx = ut , for 0 < x < 2, with the boundary conditions u(0, t) = 1
and ux (2, t) = −1.
b) 4uxx = ut , for 0 < x < 4, with the boundary conditions ux (0, t) = 2
and u(4, t) = 1.
c) (1 + t)uxx = ut , for 0 < x < 1, with the boundary conditions
u(0, t) = −1 and u(1, t) = 2.
d) uxx = ut +u, for 0 < x < 3, with the boundary conditions u(0, t) = 1
and u(3, t) = 2.
e) uxx −ux = ut , for 0 < x < 2, with the boundary conditions u(0, t) =
−1 and u(2, t) = 1.
3. Solve
∂2u
∂u 0 < x < 2,
= , for
∂x2 ∂t 0 < t,
where u(0, t) = 1, ux (2, t) = −1, and u(x, 0) = 0.
4. Solve
2 0 < x < 1,
(1 + t)∂x u = ∂t u , for
0 < t,
where u(0, t) = −1, u(1, t) = 2, and u(x, 0) = 0.
208 Chapter 7. Partial Differential Equations
5. Solve
∂2u ∂2u
0 < x < 1,
9 2 = 2 , for
∂x ∂t 0 < t,
where u(0, t) = 1, u(1, t) = −1, u(x, 0) = 1 − 2x − 7 sin(3πx), and
ut (x, 0) = 0.
∂2u
∂u 0 < x < L,
D 2 = + p(x, t) , for (7.79)
∂x ∂t 0 < t,
u(0, t) = 0, (7.80)
and
u(L, t) = 0. (7.81)
where the wn (t)’s are determined from the PDE. The expansion in (7.82)
is guaranteed from the Sine Convergence Theorem (page 189) because u is
a smooth function and it satisfies the homogeneous boundary conditions
(7.80) and (7.81).
We will also expand the forcing function p is a sine series, and write
∞
X nπx
p(x, t) = pn (t) sin , for 0 < x < L, (7.83)
L
n=1
where
L
2 nπx
Z
pn (t) = p(x, t) sin dx. (7.84)
L 0 L
Because p(x, t) is known, the pn (t)’s are known. Note that it is not
assumed that p = 0 at the endpoints, which is why the interval in (7.83)
is 0 < x < L and not 0 ≤ x ≤ L.
7.7. Inhomogeneous PDEs 209
For this to hold, the term in the square bracket must be zero. The proof
of this uses the integration formula (7.30), in exactly the same way it was
used to find the coefficients of the sine series. The conclusion is that
nπ 2
D wn (t) + wn′ (t) + pn (t) = 0,
L
or equivalently,
wn′ + κn wn = −pn , (7.86)
where nπ 2
κn = D .
L
This is a first-order linear differential equation for wn , which can be solved
using an integrating factor. The integrating factor in this case is, from
(2.18), µ = eκn t . So, from (2.21), we get the general solution of (7.86) is
Z t
−κn t κn s
wn (t) = e − pn (s)e ds + wn (0) . (7.87)
0
2 L nπx
Z
wn (0) = g(x) sin dx.
L 0 L
7.7.1 Summary
To summarize our findings, the solution of the inhomogeneous diffusion
problem (7.79)-(7.81), which satisfies the initial condition (7.88), is
∞
X nπx
u(x, t) = wn (t) sin , (7.89)
L
n=1
210 Chapter 7. Partial Differential Equations
where Z t
−κn t κn s
wn (t) = e − pn (s)e ds + wn (0) , (7.90)
0
L
2 nπx
Z
pn (t) = p(x, t) sin dx, (7.91)
L 0 L
L
2 nπx
Z
wn (0) = g(x) sin dx,
L 0 L
and κn = D(nπ/L)2 .
Example
Suppose the problem to solve is
∂2u
∂u 0 < x < 1,
4 2 = + 3 sin(2t) sin(πx) , for (7.92)
∂x ∂t 0 < t,
where
u(0, t) = 0, (7.93)
u(1, t) = 0, (7.94)
and
u(x, 0) = 0, for 0 < x < 1. (7.95)
In this problem, D = 4 and L = 1. The first step is to find the pn ’s. From
(7.91), we want
∞
X
3 sin(2t) sin(πx) = pn (t) sin(nπx).
n=1
So, p1 (t) = 3 sin 2t and all the other pn ’s are zero. Also, since g(x) = 0
then wn (0) = 0, for all n This leaves the integral in (7.90), and so
Z t Z t
κ1 s
p1 (s)e ds = 3 sin(2s)eκ1 s ds
0 0
2 + κ1 eκ1 t sin(2t) − 2eκ1 t cos(2t)
=3 .
κ21 + 4
Since κ1 = 4π 2 , then
3 h
2 −4π 2 t
i
w1 (t) = cos(2t) − 2π sin(2t) − e .
2(1 + 4π 4 )
Exercises
1. You are to find the solution of the diffusion equation (7.79), where
u(0, t) = 0, u(1, t) = 0, u(x, 0) = 0, and p(x, t) is given below. Assume
that D = 4 and L = 1.
a) p(x, t) = −4 cos(t) sin(5πx).
b) p(x, t) = e−2t sin(3πx).
c) p(x, t) = 1.
2. There is a simpler way to solve an inhomogeneous PDE when the
forcing function does not dependent on t. In this problem assume that
p(x, t) = x2 .
a) Find the steady state solution of (7.79), that satisfies (7.80) and
(7.81).
b) Letting u(x, t) = w(x) + v(x, t), where w(x) is the steady state
solution you found in part (a), find the PDE and boundary condi-
tions satisfied by v(x, t). Also, if u(x, 0) = g(x), then what is the
resulting initial condition for v(x, t)?
c) Assuming g(x) = 0, find v(x, t), and from this determine the solu-
tion of the original diffusion problem.
3. This exercise considers how to use a sine series to solve
∂2u
∂u 0 < x < 2,
2
= + 5u , for
∂x ∂t 0 < t,
where u(0, t) = 0, u(2, t) = 0, and u(x, 0) = x. This is going to be
done using the assumption in (7.82), which for this problem is
∞
X nπx
u(x, t) = wn (t) sin , for 0 ≤ x ≤ 2.
2
n=1
b) Substitute the series into the PDE and rewrite the result so it re-
sembles (7.85). From this determine the differential equation wn (t)
satisfies.
c) Find the general solution for wn (t), and from this write down the
general solution for u(x, t).
d) Use the general solution to satisfy the initial condition, and from
this determine the solution of the problem.
4. Solve
∂2u
∂u 0 < x < 3,
(1 + t) = , for
∂x2 ∂t 0 < t,
where u(0, t) = 0, u(3, t) = 0, and u(x, 0) = 1. Find the solution using
the procedure outlined in Exercise 3 (with the appropriate modifica-
tions).
∇2 u = 0, (7.96)
Later we will consider polar coordinates, and the respective formula for
∇2 will be given at that time.
It should not be a surprise that (7.96) is known as Laplace’s equa-
tion. It plays a fundamental role in applied mathematics. If you look
through a junior or senior level textbook in complex variables, fluid dy-
namics, electromagnetism, heat transfer, etc, it will appear often. As
an example, heat conduction is governed by the diffusion equation ut =
D∇2 u. So, if you want to determine the steady-state temperature distri-
bution, then you must solve (7.96).
Our goal is to find the function u(x, y) that satisfies Laplace’s equation
for (x, y) in a region, as illustrated in Figure 7.10, along with a boundary
condition u = f on the boundary of the region. To keep things simple we
will only consider simple shapes, and that means rectangular and circular.
In both cases, the method of separation of variables is used to find the
solution.
7.8. Laplace’s Equation 213
∇2 = ∇ · ∇.
∂ ∂
In Cartesian coordinates the gradient is ∇ = ∂x , ∂y , and from this you
∂2 ∂2
get that ∇2
= +
∂x2 ∂y 2
.
This can also be used to derive the formula for
2
∇ is other coordinate systems, such as polar coordinates.
where the boundary conditions are shown in Figure 7.11. So, u = 0 when
x = 0, when x = a, and when y = 0. Along the top, where y = b,
u = f (x).
The steps used in carrying out the separation of variables method
are very similar to what was done earlier. We will first find the general
solution of the PDE that satisfies the homogeneous boundary conditions.
Figure 7.11. Rectangular domain used when solving Laplace’s equation and
the corresponding boundary conditions.
214 Chapter 7. Partial Differential Equations
Assuming
u(x, y) = X(x)Y (y), (7.99)
X ′′ (x) Y ′′ (y)
=− . (7.100)
X(x) Y (y)
and
Y ′′ (y) = −λY (y) . (7.102)
This is essentially the same problem we had when solving the diffusion
and wave equations, and the general solution is
nπx
Xn (x) = cn sin , (7.105)
a
and
nπ 2
λn = − , (7.106)
a
for n = 1, 2, 3, . . .. Also, cn is an arbitrary constant.
7.8. Laplace’s Equation 215
Finding Y (y)
Yn = An enπy/a + Bn e−nπy/a .
Yn = An enπy/a − e−nπy/a
The resulting general solution, that satisfies the PDE and homogeneous
boundary conditions, is, therefore,
∞
X
u(x, y) = Xn (x)Yn (y),
n=1
or equivalently
∞
X nπy nπx
u(x, y) = cn sinh sin , (7.108)
a a
n=1
where the cn ’s are arbitrary constants. In writing this down, the constant
2An in (7.107) has been absorbed into the cn in (7.105).
2 a nπx
Z
bn = f (x) sin dx.
a 0 a
216 Chapter 7. Partial Differential Equations
With this value for cn , u(x, y) given in (7.108) is the solution of the
problem.
∂2 1 ∂ 1 ∂2
∇2 = + + .
∂r2 r ∂r r2 ∂θ2
Therefore, the problem we are solving is
1 1 0 ≤ r < a,
urr + ur + 2 uθθ = 0 , for (7.111)
r r 0 ≤ θ < 2π,
u r=a
= f (θ). (7.112)
u θ=0
=u θ=2π
and uθ θ=0
= uθ θ=2π
. (7.113)
In the vernacular of the subject, these are called periodic boundary con-
ditions. Also, note that these boundary conditions are homogeneous be-
cause u = 0 satisfies both of them. Finally, if (7.113) hold then ur is also
continuous in the domain.
Assuming
u = R(r)Θ(θ), (7.114)
and then substituting this into Laplace’s equation (7.111) gives us
and
Θ ′′ (θ) = −λΘ(θ) . (7.117)
From (7.113) we also must have that
As usual, the first problem to solve is the one involving the homogeneous
boundary conditions, which means solving (7.117) and (7.118).
λn = n2 , for n = 1, 2, 3, . . . , (7.119)
and
Θn = An cos(nθ) + Bn sin(nθ). (7.120)
Finding R(r)
λ = n2 : Now (7.116) is
r2 R ′′ + rR ′ = n2 R.
This is also an Euler equation, and the general solution is Rn =
An rn + B n r−n . Because the solution is bounded we require B n = 0.
or equivalently
∞
1 X
rn an cos(nθ) + bn sin(nθ) .
u = a0 + (7.121)
2
n=1
The coefficients in this formula are written in a form similar to what was
used earlier for a sine and cosine series. So, for example, we have written
R0 Θ0 = A0 A0 = 21 a0 .
The an ’s and bn ’s are determined in the same way as for a sine and cosine
series. For example, to determine a7 you multiply the above equation by
cos(7θ), integrate for 0 ≤ θ ≤ 2π, and use orthogonality conditions such
as given in (7.30) and (7.39). The resulting formulas obtained in this way
are Z 2π
1
an = f (θ) cos(nθ)dθ,
πan 0
and
2π
1
Z
bn = f (θ) sin(nθ)dθ.
πan 0
∇2 u = 0 , for x2 + y 2 < 1,
So, b4 = 3, and the other coefficients are zero. Therefore, the solu-
tion is
u = 3r4 sin(4θ).
The resulting solution is shown in Figure 7.14.
Exercises
1. You are to find the solution of the problem shown in Figure 7.11.
Assume that a = 1 and b = 2. Note that you should be able to answer
this question without using integration.
a) f (x) = 5 sin(2πx)
b) f (x) = −3 sin(12πx)
c) f (x) = sin(πx) − 7 sin(8πx)
d) f (x) = −3 sin(4πx) − sin(7πx) + 6 sin(20πx)
2. You are to find the solution of the problem shown in Figure 7.13.
Assume that a = 2. Note that you should be able to answer this
question without using integration.
a) f (θ) = 4 cos(3θ)
b) f (θ) = 1 − 3 sin(15θ)
c) f (θ) = sin(θ) + 3 cos(5θ)
d) f (θ) = 4 − 2 sin(5θ) − 4 sin(9θ) + 8 cos(14θ)
The following is a brief summary of the rules of matrix and vector algebra
in two dimensions.
Examples:
1 −3 −2 1 −2 1 0 2 −2
+ = + =
2 4 6 −3 4 −7 3 1 7
Examples:
1 4 1 0 −4 0
4 = −4 =
−3 −12 −2 3 8 −12
4 1 6 0 2 0
=4 =3
−8 −2 3 −12 1 −4
223
224 Appendix A. Matrix Algebra: Summary
a b
=I means that a = 1, b = 0, c = 0, d = 1
c d
Examples:
1 −2 1 −3
=
−3 4 2 3
1 −1 −1 1 −2 1 −7
3 − =3 − =
0 2 1 5 2 5 1
Examples:
t + t3 1 + 3t2
d
=
dt sin t cos t
d 1 5t 1 5t 1
e = 5e = 5 e5t
dt −2 −2 −2
Appendix B
Answers
Chapter 1
Section 1.2, pg 4
2a) r = −2 2g) none 3c) r = 1/3, c = 3
2b) r = 1/3 2h) r = 0 3d) r = 1, c = −1
2c) none 2i) none 3e) r = −2/5, c = −7
2d) r = 0, −4 2j) none 3f) r = −4, c = 3
2e) r = −3, 1/2 3a) r = −2, c = 1
2f) r = 2 3b) r = −1, c = −1
Chapter 2
Section 2.1, pg 12
2
1a) y = (9t + c)−1/3 and y = 0 2e) y (t) = ln 1/2 t e+2
e
1b) y = ±(2e−t + c)−1/2 and y = 0 √
2f) y (t) = −2 + 4 + 2 t
1c) y = −1/(cos
p t + c) and y = 0 2g) y (t) = 2 arctan (1 + t)
1d) y = 3 ± t2 /2 + c −1
2h) y (t) = 5 1 + 4 e5 t
1e) y = − ln( 12 t2 + 2t + c)
2i) y (t) = 21 ln e−2 t + e2 − 1 + t
1f) y = − 31 ln[3 ln(t + 1) + c] √
1g) y = − 41 ln(2e2t + c) 2j) y (t) = ln t/2 + 1/2 t2 + 4
1h) y = − ln12 ln[t ln(2) + c] 3a) q (r) = − √141r+1
1i) y = 31 [−1±(6t+c)−1/2 ], y = − 13 3c) h (τ ) = −2 + 4 eτ /3
−1
1j) y = −2 − 1/(t + c) and y = −2 3d) h (x) = 6 2 + e3 x
1k) y = 3 − 2/(t + c) and y = 3 3e) z (r) = 6 (1 + 6 ln ((1 + er)/2))
−1
1l) y = tan(t/3 + c) 3f) w (τ ) = 1/2 ln 1/8 τ 4 + 1
1m) y = ln[tan(t2 /2 + c)] 2
t 3g) r (θ) = 2 (θ +√1)
1n) y = ln(ce
√ − 1)
3h) r (θ) = −1 + 2 θ2 + 1
1o) y = ± cet2 − 1
4a) y − ln(1 + y) = t + 1 − ln 2
2a) y (t) = 5 √1501 t+1
4b) 15 t = y 5 + 5 y + 6
2c) y (t) = 4 + 7 t 4c) y + ln(1 + y) = t + 5 + ln(6)
2d) y (t) = (1 + ln (4 + et ) − ln (5))
−1
4c) y − e−y = t + 2 − e−2
Section 2.2, pg 18
225
226 Appendix B. Answers
Section 2.3, pg 28
ln(2)
1c) ln(4/3) days 8d) 792 + 968e−20/11 ft
2d) either 40 or 39 BC 8e) −22 fps
3c) 50 ln(10) min 9e) m/c − cL/A √
11a) P = N4 4 + z − 8z + z 2 , z = e−rt
4c) 5(1 − e−6 ) g
5b) 104 (1 − e−1 ) kg 9−e−2t
12a) P = 250 3−e−2t
25 11
6c) 324000
11 1 − 27 lbs 13d) 5 ln(64/39)
ln 2 min
7a) v = −20 + 120e−t/2 m/s 14c) k = (4 − 21/4 )/5
1/4
7c) 40(5 − ln 6) m 15b) 120 ln(42/37)
ln(4/3) min
8a) v = −(176/c)(1 − e−2ct/11 ) fps
15c) 120 ln(93/74)
ln(4/3) min
Section 2.4, pg 38
us=unstable; as=asymptotically stable
1a) as 2c) y = ±2, as; y = 0, us
1b) us 2d) y = − ln 2, as
1c) as 2e) y = −2, as; y = 2, us
1d) us 2f) y = 0, as; y = ln 3, us
2a) y = 1, us; y = −2, as 9c) 750
2b) y = ±1, us; y = 0, as
Chapter 3
Section 3.5, pg 50
1a) −7, 1 3d) y (t) = c1 e−t/2 + c2 et/2
1b) −2,√ 2 3e) y = c1 + c2 t
1c) 21 e2 3, 12 e2 3f) y (t) = c1 e3 t + c2 e3 t t
√
1d) 1 + 2e2 3, 1 + 2e 2 3g) y (t) = c1 e−t/2 + c2 e−t/2 t
1
√ 2 1
√
1e) 2 ( 3 − 1)e , 2 ( 3 + 1)e2 3h) y = c1 sin (t/2) + c2 cos (t/2)
1f) −e12 , 0 3i) y (t) = c1 et sin (t) + c2 et cos (t)
3a) y (t) = c1 e−2 t + c2 et 3j) y = e−t (c1 sin (2 t)+c2 cos (2 t))
3b) y (t) = c1 e−2 t + c2 et/2 4a) y (t) = −1/3 e2 t + 1/3 e−t
3c) y (t) = c1 + c2 e−3 t 4b) y (t) = −4/5 et/2 − 1/5 e−2 t
227
4c) y (t) = −4/3 + 1/3 e−3 t 4i) y = −e−t sin (2 t) − e−t cos (2 t)
4d) y (t) = 4 − 5 et/5 √ 4j) y = 2 et/2 cos (t/3)
3
4e) y = 3 exp(−(1/3) 3t) 6a) (1 + t)
2
4f) y = 5 exp(−(1/2)t) 6b) cos t + 6 t
4g) y (t) = −e−t − e−t t 6c) t + 2
4h) y (t) = −1/3 sin (3 t) − cos (3 t)
Section 3.8, pg 58
1a) y (t) = e3 t c2 + e−2 t c1 − et
2
1b) y = c1 e−2t + c2 e−t + −π sin(πt)−3π cos(πt)+2 sin(πt)
π 4 +5π 2 +4
1c) y (t) = e−5 t c2 + et c1 − 2/5 t2 − 16 t
25 − 125
84
3 sin(2 t) 15 cos(2 t) 15
1d) y (t) = 5 et/5 c1 − 202 − 101 − 101 + 1/6 e−t + c2
15 t2
2t
1e) y (t) = e c2 + e −t/3
c1 − 1/2 t + 4 − 4 + 535
3 89 t
8
1f) y (t) = e−t/4 c2 + c1 et/2 + 4 cos(2 221
t)
− 33 sin(2 t)
221 − 4
1g) y (t) = sin (2 t) c2 + cos (2 t) c1 + 1/25 (5 t − 2) et
1h) y (t) = c2 e−t + c1 e6 t + +1/9 (3 t − 1) cos (3 t) + 1/15 (−5 t − 2) sin (3 t)
1i) y (t) = sin (2 t) et c2 + cos (2 t) et c1 + t2 + 4/5 t + 18 25
1j) y (t) = e−t sin (3 t) c2 + e−t cos (3 t) c1 + 1/10 + 3/13 et
1k) y (t) = 1/3 e3 t c1 − 1/9 t2 − 1/9 t3 − 1/12 t4 + 52 t
27 + c2
2/3 t −2 t 3 t −2 t
1l) y (t) = e c2 + e
−t
c1 − 1/2 e e + 3/8 e
1m) y (t) = e4 t sin (t) c2 + e4 t cos (t) c1 − 1/2 e4 t t cos (t)
1n) y (t) = e6 t c2 + e−t c1 − 15 cos(t+7)
74 + 21 sin(t+7)
74
1o) y (t) = −e−2 t c1 − 3 sin(2 40
t)
+ 1/4 + 1/40 cos (2 t) + e−t c2
1p) y (t) = −4 e−t/4 c1 − 2 sin(2 65
t)
− cos(2 t)
260 + c2
t −2 t
2a) y (t) = e − 1/4 e − 3/2 t − 3/4
2b) y (t) = −1/8 + 9 cos(2 8
t)
+ 1/4 t2
t
2c) y (t) = −1/2 e − 1/2 sin (t) + 1/2 cos (t) + 1
−3 t
2d) y (t) = − 2 e27 + 1/3 t2 − 2/9 t + 29 27
e−2 t
2e) y (t) = 19 16 + 11/4 e−2 t t − 3/16 e2 t
2f) y (t) = 1/2 e−t/2 + 3/4 et/2 − 1 + 1/4 (−t − 1) e−t/2
2g) y (t) = −1/9 sin (3 t) + 1/3 cos (3 t) t
2h) y (t) = −1/2 e−t sin (2 t) − 5/4 e−t cos (2 t) + 1/4 e−t
t/2 t/2
2i) y (t) = 51 e 13 sin(t/2)
+ 21 e 13 cos(t/2)
− 12 sin(3
13
t)
− 34 cos(3
13
t)
t 6t
4a) y (t) = −2/5 e + e c
4b) y (t) = e−2/3 t c + −3 π cos(π t)+2 sin(π t)
9 π 2 +4
4c) y (t) = 2/3 t − 2/9 + e−3 t c
4d) y (t) = −3 t − 15 − 1/6 e−t + et/5 c
4e) y (t) = − 101
t cos (2 t) − 1/25 cos (2 t) − 1/5 t sin (2 t) − 3 sin(2
100
t)
+ e4 t c
6t
4f) y (t) = −1/7 e t − 1/49 e − 1/3 + e c
−t −t
4g) y (t) = −1/10 e−t cos (t) + 3/10 sin (t) e−t + e−2/3 t c
4h) y (t) = −1/4 cos (2 t + 5) + 1/4 sin (2 t + 5) + e2 t c
Section 3.9, pg 63
1a) y = −2 e−2 t + 2 et/2 − 5 e−2 t t
1b) y = 3 + (−3 cos (t) + 3 sin (t)) et
Rt Rt
1c) y = −e−2 t 0 ln (1 + s) e2 s ds + et 0 ln (1 + s) e−s ds
t
1d) y = t/3 + 2/15 t5/2 + e−3 t 0 −1/3 s3/2 + 1 e3 s ds
R
228 Appendix B. Answers
R t −s/5
1e) y = −2 ln (t + 1) + et/5 0 2 e1+s ds
t Rt
1f) − 41 e−t/2 0 sin s2 + 1 es/2 ds + 14 et/2 0 sin s2 + 1 e−s/2 ds
R
3a) 2 t (−t + et − 1)
3b) 1/2 (t − 1) e2 t + 1/2 + t/2
3c) 4 t5/2 √
4b) 1/2 sin (t) t
Section 3.10, pg 73
√ −2t √
2b) u = 14 cos(8t − π) 1
7b) u = − 90 15e sin(2 15t)
2e) 1 √ √ 7d) e−τ /24, τ = z( 3π2 −Arctan(z)),
2 5π
√
3b) u = 3 3 cos 2 3t − 6 z = 1/ 15
√ √
3f) 3π/18√ 8d) u = 2e−3t cos(t − π/4)
1
2 cos 10t − 7π
4b) u = 20 12b) u (t) = 3/4 sin (16 t) t
√ 4
4e) 5(2 + 2); 3π/40 13b) u (t) = 5 sin(3
12
t)t
Section 3.11, pg 77
1a) y (x) = c1 x2 + c2 x2 ln (x)
1b) y (x) = c1 x3 sin √ (ln (x)) + c2 x3 cos (ln (x))
c
1c) y (x) = √1x + c2 3 x
√ √ √ √
1d) y (x) = c1 x sin 1/2 3 ln (x) + c2 x cos 1/2 3 ln (x)
1e) y (x) = c1 x2 sin (3 ln (x)) + c2 x2 cos (3 ln (x))
1f) y (x) = cx1 + xc2/5
2
Chapter 4
Section 4.3, pg 91
2a) 3, −2 2b) 5, −1 3a) 2 ± 2i 3b) −1 ± 2i
Section "
4.5, pg 98
c 1 e−3 t + c 2 e2 t
#
1a)
−1/3 c 1 e−3 t + 1/2 c 2 e2 t
c 1 e−t/2 + c 2 et/2
" #
1b)
−2 c 1 e−t/2 + 2 c 2 et/2
c 1 + c 2 e5 t
" #
1c)
5t
" −2 c 1 + 3 c2 t2 e #
−c 2 e
1d)
c 1 e2 t + c 2 e2 t t
c 1 e−2 t
" #
1e)
−2 t
" c 2e #
c 1 sin (3 t) + c 2 cos (3 t)
1f)
−1/3 c 1 cos (3 t) + 1/3 c 2 sin (3 t)
229
c 1 sin (t/2) + c 2 cos (t/2)
1g) c 1 (−1/5 sin (t/2) + 1/10 cos (t/2))
2t
c 1 te + c 2 2et + c 3 e t
−t −t
c 1 e + 3c 2 e + 2c 3 e t
4b)
c 2 e2 t
2t t
c 2 e +c 3 e
2 c 1 et − c 2 e−2 t
4c)
3 c 1 et
t −2 t
7c 1 e + 4c 2 e +c 3 e −t
c 1 e−t
√ √
4d)
2 c 2 e 5t + 2 c 3 e− 5t
√ √
− 5t √
5t
√
−c 3 5+1 e +c 2 e 5−1
Section 4.7, pg 110
us=unstable; as=asymptotically stable; ns=neutrally stable
1a) us 1d) us 1g) as 3a) us 3d) us
1b) as 1e) as 1h) ns 3b) as
1c) us 1f) us 1i) ns 3c) as
Chapter 5
Section 5.1, pg 115
2a) (u, v) = (1/2, 0), (1/3, 1/4) 2e) (S, I) = (5, 0), (1, 2)
2b) (u, v) = (0, 0), (−1, 1) 2f) (s, c) = (−1, 1)
2c) (u, v) = (1/4, 4) 2g) (x, y) = (0, 0)
2d) (S, P ) = (1, 1), (0, 0), (2, 0) 2h) (x, y) = (0, 2), (0, −1), (2, 0)
Section 5.2, pg 126
1a) (u, v) = (1/2, 0) as; (1/3, 1/4) as
1b) (u, v) = (0, 0) us
1c) (u, v) = (1, −1) us
1d) (u, v) = (1/4, 4) as
230 Appendix B. Answers
Chapter 6
Section 6.1, pg 149
e−1−s
+ 3 1−es
−1 −s
1a) − (s − 5) 2d) 1+s
1b) 4+3s2
s
2e) 5 s − 1−2se2
−1 −3 s
−1
1c) 2 s−2 + 4 (1 + s) 2f) 2 e
−12 s
−e−15 s
1 s
1d) −9 (s+2)(s−7) 3a) s2s+9
1e) 8 s −3 −1
2 3b) 28 s2 + 49
1f) 9 s −6s3
s+2
3c) (s+1)s+1
2
−2 +π 2
1g) (1 + s) 2
s +3
−3 cos(4)π−sin(4)s 3d) 2 (s2 +9)(s 2 +1)
1h) −2 9 π 2 +s2 s
5 s2 −8 s+4
4a) 6 (s2 +9)2
1i) (s−2)(s2 +4) 2
−2 s 4b) 6 (ss2 +49)
−49
2
2a) 5 e s
−2 s s(s2 −3)
2b) 1−es2 4c) 2 (s2 +1)3
−2 s s+2
2c) − 1−2 e s +e 4d) 10
−s
2
((s+2)2 +25)
6s
1g) 1−e +3 es −e
−s −2 s −3 s
2h) cosh 2t − cosh (t)
−2 s −4 s −8 s 2i) sin (2 t) − sin (3 t)
1h) e +e s −e
231
2j) 5 t − 3 e2 t 1
− es
−s
4a) 1−e−2 s s
−1
4 e2 t
" # " #
s 1/4
1
1e) 3b) s2 −1/4
−e2 t − 4 e2 t t 1 s
" 1 t
− 2 e sin (2 t) + 4 et cos (2 t)
# " #
s−2 0
1
1f) 3c) (s−2)2
−et cos t
" (2 t) − 8 e# sin (2 t) "−1 s − 2 #
s 6 s − 1 −4
1 1
3a) s2 +s−6 3d) s2 −2 s+5
1 s+1 1 s−1
Chapter 7
Section 7.2, pg 179
e2 x e−2 x
1a) u (x) = − e−4 −e4 + e−4 −e4
1b) u (x) = − sin(x)
sin(1)
√
e−x/2 sin(1/2 3x)
1c) u (x) = e−1/2 sin(1/2
√
3)
e−x (−1+e2 ) ex (−1+e−2 )
1d) u (x) = −5 e−2 −e2 + 5 e−2 −e2 − 5
e−x
−1
1e) u (x) = 1/2 x2 + 1/2 e−1 −1 − x − 1/2 e−1 − 1
un = bn sin π2 (2n − 1)x , with λn = −[ π2 (2n 2
3a) − 1)]
nπ
3b) u0 = b0 , with λ0 = 0; and un = bn cos 4 x , with λn = −( nπ 4 )
2
p
3c) u = be−λx/2 sin(πx/4), with λ = ± 4 − (π/2)2
3d) un = bn e−x/2 sin(nπx), with λn = − 14 − (nπ)2
3e) u0 = b0 , λ0 = 0; un = an sin(2πnx) + bn cos(2πnx), with λn = 4π 2 n2
Section 7.3, pg 186
2
1a) −4 e−75 π t sin (5 π x)
2
1b) 6 e−363 π t sin (11 π x)
2 2 2
1c) e−3 π t sin (π x) + 8 e−48 π t sin (4 π x) − 10 e−147 π t sin (7 π x)
2 2 2
1d) −e−27 π t sin (3 π x) + 7 e−192 π t sin (8 π x) + 2 e−675 π t sin (15 π x)
2 2
1e) 2 e−27 π t sin (3 π x) + 2 e−3 π t sin (π x)
n −n2 π 2 t
2a) n=1 −2 (−1+(−1) )e n π sin(1/2 n π x)
P∞
n −n2 π 2 t
2b) n=1 −4 (−1+2 (−1) )e n π sin(1/2 n π x)
P∞
−π 2 t n −n2 π 2 t
2c) −4/3 e sin(1/2 π x)
+ n=3 −2 n (−1+(−1) )e sin(1/2 n π x)
P∞
π π (n2 −4)
n −n2 π 2 t
3) n=1 30 (−1) e sin(1/3 n π x)
P∞
nπ
P∞ (2 cos(1/4 n π)−2 cos(3/4 n π))e−5/2 n2 π2 t sin(1/2 n π x)
4) n=1 nπ
2
5a) n=1 bn e−kn t sin(kn x), kn = π(2n − 1)/2
P∞
2
5b) n=1 bn e−4kn t cos(kn x), kn = π(2n − 1)/2
P∞
2 2
5c) n=1 bn e−kn (t+t /2) sin(kn x), kn = nπ
P∞
2
5d) n=1 bn e−kn t+e sin(kn x), kn = nπ
P∞ −t
2 2
6d) u = −24e(−k3 t+e −1) sin(k3 x) − 12e(−k15 t+e −1) sin(k15 x)
−t −t
∞
X n sin (1/2 n π x)
4c) 4
n=1
π (n2 − 4)
n odd
n
6 (−1) −8 cos(1/4 n π)
+ 2 n1π sin (1/2 n π x)
P∞
4d) n=1 nπ
P∞
4e) n=1 −2 n π cos(1/2nn2π)−4 π2
sin(1/2 n π)
+ 4 n1π sin (1/2 n π x)
P∞ n
4h) n=1 −6 (−1) −2 cos(1/6 nπ
n π)+6 cos(2/3 n π)
+ 2 1
n π sin (1/2 n π x)
P∞ ((−1)n −1) cos(1/2 n π x)
5a) 1 + n=1 4 n2 π 2
2
P∞ ((−1)n e2 −1) cos(1/2 n π x)
5b) −1/2 + 1/2 e + n=1 4 n2 π 2 +4
5c) cos(πx)
5d) −2 + n=1 8 sin(1/4 n π)ncos(1/2 n π x)
P∞
π
5e) 3/4 + n=1 2 n π sin(1/2 nnπ)−4 cos(1/2 n π) 1
P∞
2 π2 + 4 n2 π 2 cos (1/2 n π x)
P∞ (2 sin(1/6 n π)−6 sin(2/3 n π)) cos(1/2 n π x)
5h) 7/6 + n=1 nπ
n
8a) 1/3 + n=1 4 (−1) ncos(n π x)
P∞
2 π2
n
9a) n=1 −2 (−1) nsin(n π x)
P∞
π
Section 7.5, pg 204
1a) cos(12πt) sin(3πx)
1
1b) − 16π sin(32πt) sin(8πx)
1c) − cos (4 π t) sin (π x) + 4 cos (12 π t) sin (3 π x) − 3 sin(20 π20t)πsin(5 π x)
1d) 5 cos (28 π t) sin (7 π x) + sin(32 π 16π t) sin(8 π x)
+ sin(48 π t)16π
sin(12 π x)
sin(32 π t) sin(8 π x)
1e) cos (12 π t) sin (3 π x) + cos (4 π t) sin (π x) − 16π
1f) 3Pcos (8 π t) sin (2 π x) − sin(36 π 24π t) sin(9 π x)
+ 3 sin(20 π40t)πsin(5 π x)
∞
2a) n=1 an cos(kn t) + bn sin(kn t) sin(k n x), kn = (2n − 1)π/2
P∞
2b) n=1 an cos(2kn t) + bn sin(2kn t) cos(kn x), kn = (2n − 1)π/2
P∞
2c) a + bt + n=1 an cos(nπt) + bn sin(nπt) An cos(2nπx) + Bn sin(2nπx)
√
2d) e−t/2 n=1 an cos(ωn t) + bn sin(ωn t) sin(nπx) , ωn = 4n2 π 2 − 1/2
P ∞
n
3) n=1 −2 (−1+(−1) ) sin(2 n π t) sin(n π x)
P∞
n4 π 4
Section 7.6, pg 207
n −4 n2 π 2 t
1a) 1 − 2 x + n=1 −2 (1+(−1) )e n π sin(n π x)
P∞
n −4 n2 π 2 t
1b) 2 − 7 x + n=1 −14 (−1) e n π sin(n π x)
P∞
−4 n2 π 2 t
1c) −4 + 5 x + n=1 8 e sin(n π x)
P∞
nπ
2a) 1 − x
2b) −7 + 2x
2c) −1 + 3x
2d) Aex + Be−x , where A = (2 − e−3 )/(e3 − e−3 ), B = (e3 − 2)/(e3 − e−3 )
2e) A + Bex , where A = (1 + e2 )/(1 − e2 ), B = 2/(e2 − 1)
n −n π (t+1/2 t )
2 2 2
n −(5+1/4 π n )t
2
235
236 BIBLIOGRAPHY
237
238 Index